Science and the Doctrine of Creation: The Approaches of Ten Modern Theologians 0830852808, 9780830852802

Can Christians take seriously the claims of modern science without compromising their theological integrity? Can theolog

305 42 6MB

English Pages 264 [268] Year 2021

Report DMCA / Copyright

DOWNLOAD PDF FILE

Recommend Papers

Science and the Doctrine of Creation: The Approaches of Ten Modern Theologians
 0830852808, 9780830852802

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

SCIENCE AND THE

D O C T R I N E OF C R E AT ION THE APPROACHES OF TEN MODERN THEOLOGIANS

EDITED BY

GEOF F R E Y H . F U LKE RSON A N D JOE L T HOMA S C HOP P AFTERWORD BY ALISTER E. M C GRATH

InterVarsity Press P.O. Box 1400, Downers Grove, IL 60515-1426 ivpress.com [email protected] ©2021 by Geoffrey H. Fulkerson and Joel T. Chopp All rights reserved. No part of this book may be reproduced in any form without written permission from InterVarsity Press. InterVarsity Press® is the book-publishing division of InterVarsity Christian Fellowship/USA®, a movement of students and faculty active on campus at hundreds of universities, colleges, and schools of nursing in the United States of America, and a member movement of the International Fellowship of Evangelical Students. For information about local and regional activities, visit intervarsity.org. All Scripture quotations, unless otherwise indicated, are taken from The Holy Bible, New International Version®, NIV ®. Copyright © 1973, 1978, 1984, 2011 by Biblica, Inc.™ Used by permission of Zondervan. All rights reserved worldwide. www.zondervan.com. The “NIV” and “New International Version” are trademarks registered in the United States Patent and Trademark Office by Biblica, Inc.™ The publisher can’t verify the accuracy of website hyperlinks beyond the date of print publication. Cover design and image composite: Bradley Joiner Image: eclipse: scott-szarapka-y07ClgzcVmc-unsplash ISBN 978-0-8308-2675-9 (digital) ISBN 978-0-8308-5280-2 (print)

CONTENTS Acknowledgments

vii

Introduction by Geoffrey H. Fulkerson and Joel Thomas Chopp

1

1 William Burt Pope (1822–1903)

13

2 Abraham Kuyper (1837–1920)

35

3 B. B. Warfield (1851–1921)

59

4 Rudolf Bultmann (1884–1976)

84

5 Karl Barth (1886–1968)

99

Primary and Secondary Creation by Fred Sanders Enlightenment, Science, Worldview, and the Christian Mind by Craig Bartholomew Evolution, Human Origins, and the Development of Theology by Bradley J. Gundlach Myth, Science, and Hermeneutics by Joshua W. Jipp The Doctrine of Creation and the World of Science by Katherine Sonderegger

6 T. F. Torrance (1913–2007)

121

7 Jürgen Moltmann (1926–)

148

8 Wolfhart Pannenberg (1928–2014)

170

9 Robert Jenson (1930–2017)

192

Christ the Key to Creation and Theological Science by Kevin J. Vanhoozer The Environment of Science and Theology by Stephen N. Williams Nature, Contingency, and the Spirit by Christoph Schwöbel History’s God by Stephen John Wright

10 Colin E. Gunton (1941–2003)

213

Afterword by Alister E. McGrath

241

The Triune God, Scientific Endeavor, and God’s Creation Project by Murray A. Rae

Contributors | 249

General Index | 253

Scripture Index | 256

Praise for Science and the Doctrine of Creation

257

About the Author

259

More Titles from InterVarsity Press

260

AC K N OW L E D G M E N T S

T

he contents of this book have their origin in a multivenue event called “A Modern Creature: Science and the Doctrine of Creation in Modern Theology,” which was primarily a two-day conference hosted by the Carl F. H. Henry Center for Theological Understanding at Trinity Evangelical Divinity School. The Henry Center also published a series of shorter reflections in tandem with the conference hosted on Sapientia, our digital periodical. All of this was made possible through the generous support of a grant from the Templeton Religion Trust. We are grateful to all our authors, not only for their original contributions but also their revised and expanded versions published here. We are also thankful to Matt Wiley for his extensive editorial assistance, TaeJung Eric Kim for his careful reading and systematic development of the index, and Peter Highley for his insightful comments on a draft of the manuscript. Tom McCall, previous director of the Henry Center, both supported us in this project and encouraged us along in our various academic endeavors. And finally, this project would not have been possible without the support and editorial guidance of Jon Boyd, whose patience and encouragement played no small role in ensuring that it would see the light of day. Warm thanks are also due to the entire team at IVP for their excellent work in bringing this volume to print.

I N T R O D U C T IO N Geoffrey H . Fu lkerson and Joe l Tho m as C hopp

S

cience and the Doctrine of Creation examines the reception of the natural sciences among Protestant theologians in the modern era, particularly with respect to the doctrine of creation. While there is no shortage of handbooks, introductions, and the like on the relation between science and religion in the abstract, comparatively little research is available on how leading theologians actually engaged particular theories or developments in the natural sciences. This volume is not primarily a constructive work in dogmatic theology, giving a normative account of how Christians should engage scientific developments, nor is it focused on any particular set of scientific questions and challenges. Rather, it is a study in the history of Christian thought intended to introduce the unique contributions of Protestant theologians working in an age increasingly shaped by the assumptions, claims, and methods of the natural sciences. While much of the contemporary discussion at the intersection of science and theology focuses on the doctrine of creation, it also tends to prioritize questions related to human and universal origins, which traditionally have been only one facet of the doctrine. Drawing on a broad stream of Christian reflection, John Webster identified three topics traditionally treated within the doctrine of creation: the identity of the Creator, the act of creating, and the natures and ends of created things.1 When one of these elements receives attention to the neglect of the others, the doctrine itself suffers from atrophy and disorder. For example, the John Webster, “Love Is Also a Lover of Life: Creatio Ex Nihilo and Creaturely Goodness,” Modern Theology 29, no. 2 (April 2013): 157.

1

2

Geoffrey H. Fulkerson and Joel Thomas Chopp

Christian affirmation of the primordial goodness of the created order is understood to be grounded in the nature of God, not merely an optional description tacked on to the end of God’s act of creation in the Genesis narrative. Even the classical doctrine of creation ex nihilo, which figures in several of the following chapters, is about both the original act of creation and God’s radical perfection that conditions the nature of God’s relation to what is other than himself. When creation moves from that which is created (an original act) to that which is a creature, which remains stable and open to continuing divine interaction, the world itself begins to appear as a stable structure and reliable object of inquiry, however much we take it for granted. Scripture often makes this point by invoking the “pillars” and “foundations” on which God created the earth. Far from a naive cosmology, this is a transparent architectural metaphor, tied to the strength and stability of God’s creation. God made the earth, and God alone sustains it (or shakes it). Moreover, these foundations refer not only to material things like trees and animals but also to forces (e.g., gravity) as well as the structure of intelligibility—its unity, meaning, and purpose—such that we can invoke ideas like “world” and “history.” With this third sense of creation, we also touch on what is variously called natural theology or a theology of nature. Our categories refer intelligibly to things because God made it so, which is why even those without Scripture can live in and think truthfully about the world. But it also raises a question: How are we to think rightly about the world that God created? We tend to raise this question either in relation to issues in the natural sciences, such as evolution, genetics, and chance, or in the human sciences with categories like power, ideology, and identity. However different the questions and categories, this is one facet of the age-old question of the relationship between philosophy and theology. In this sense, Darwin’s reflections on adaption or Marx’s reflections on ideology are not fundamentally different from Aristotle’s reflections on substance, nor is the challenge that each presents to the church: whether we should “plunder the Egyptians” and, if so, how we might do so without turning their gold into new idols. This raises fundamental questions about the nature of reality (ontology, cosmology) and our access to it (epistemology),

Introduction

3

which are shared by theology, metaphysics, science, and the humanities alike. It also raises questions about the conceptual schemes that we employ and how they relate to our theological understanding. The chapters that follow engage creation at all three levels. The overall picture that emerges from the volume illustrates the complex way that creation is what John Webster has called a “distributed doctrine.” Rather than being a topic that we can expound by isolation and abstraction, the doctrine of creation appears throughout the dogmatic corpus, from theology proper to anthropology and eschatology. While creation might appear in these other doctrines, as Webster helps us see, it is equally true that these other doctrines appear in our exposition of the doctrine of creation. It is the distributed nature of the doctrine that especially interests us in this volume. What ancillary doctrine do theologians juxtapose alongside their doctrine of creation, how does this inform their overall picture of creation, and what insights might it provide in our engagement with the various questions and concerns that confront the church, particularly in relation to the natural sciences? Structure of the Volume We address these questions by examining the approaches of ten modern theologians. For those unfamiliar with these figures, we have provided a few aids. Each chapter begins with a brief biography of the theologian to situate them within their historical context and concludes with a list of recommended works for further reading. These theologians do not represent anything like a uniform approach to theological engagement with the sciences. Indeed, in many cases they might disagree with one another or even use the same words in significantly equivocal ways. Because of this, we have deliberately left some of the key concepts undefined. Our rationale for doing so is partly of necessity—what counts as “science” is a territory with variegated terrain and a notoriously porous boundary— but also partly heuristic in nature. When the chapters are read as a whole and placed in dialogue with one another, a conversation among Protestant theologians can be seen to emerge, despite the disparate nature of the scientific and theological issues being addressed. By wrestling across the chapters, readers will (we hope) find themselves grappling with wider

4

Geoffrey H. Fulkerson and Joel Thomas Chopp

tensions that remain inherent in the current discussions at the intersection of theology and science. As for the content of the ten chapters, we begin with theologians from two Protestant traditions that shaped nineteenth-century North America, Methodist theologian William Burt Pope (Fred Sanders), and the Reformed Princetonian, B. B. Warfield (Bradley Gundlach), and one influential European, Abraham Kuyper (Craig Bartholomew). Fred Sanders opens our collection with an account of William Burt Pope’s doctrine of creation, focusing on his three-volume Compendium of Christian Theology. Pope’s doctrine of creation programmatically distinguishes between the divine act of primary creation (God calling all things into existence) and the divine work of secondary creation (the formation of an ordered universe). The former is the realm of metaphysical inquiry and apologetic argumentation; Scripture chooses to say little about it, and science can say nothing in principle. Secondary creation, on the other hand, is the domain addressed by both the biblical account and also scientific investigation. Along the way Sanders considers Pope’s exegesis of the creation narrative in Genesis and how the distinction between primary and secondary creation operates in his judgments regarding evolution and human origins. Sanders’s chapter introduces Pope’s overall theological project and then examines how he deploys the distinction between primary and secondary creation in order to clarify and conserve the main lines of Christian dogmatics. The chapter concludes with a rich consideration of the usefulness of Pope’s distinction for a Christian doctrine of creation today. Craig Bartholomew’s account of Abraham Kuyper and Bradley Gundlach’s chapter on B. B. Warfield provides an interesting juxtaposition with one another. Kuyper and Warfield were contemporaries and indeed friends. Kuyper’s most substantive work on science and especially evolution appears in his Princeton Stone Lectures (later published as Lectures on Calvinism), which Warfield invited him to present. While Kuyper and Warfield share similar material positions, both regarding science and theology, they also have quite different intuitions about how theology was to interact with the science of their day and, correspondingly, quite different accounts of the relationship not only

Introduction

5

between science and theology but also between each of these and the wider culture.2 In Bartholomew’s chapter on Kuyper, two important threads rise to the surface that are, in fact, in tension with one another. On the one hand, creation, common grace, and the image of God all point to the shared yet independent nature of science(s). On the other hand, the fact that reality is ultimately grounded in God and that corrupt humanity cannot properly approach God means that all nonregenerate thinking is ultimately flawed. Moreover, when combined at the level of cultural analysis, these twin themes helped Kuyper develop his approach to “worldview analysis.” The theologian’s task is not to respond to this or that discrete issue or body of scientific analysis but rather to attend to the various claims made at the comprehensive level. After laying out this Kuyperian account of “science,” Bartholomew shows how Kuyper applied it to his careful reflections on the many faces of evolutionary theory as it appeared all around him. Gundlach’s chapter examines Warfield’s complex views on theistic evolution. Warfield has been claimed by evangelicals as a defender of theistic evolution, an opponent of evolutionary theory, and a compromiser whose views on human origins undermined the gospel. Through careful engagement with Warfield’s views as they developed throughout his career, Gundlach shows that his position defies easy categorization. More germane to the volume at hand, Warfield demonstrated a form of theological engagement that humbly, thoughtfully, and openly engaged the prevailing scientific theories of his day while also remaining grounded in Scripture and versed in the tradition. While remaining “agnostic” to any particular scientific theory, Warfield is in search of some theologically appropriate form of evolution. Gundlach speaks of “Warfield’s pervasive developmentalism,” evident not only in evolutionary theory but also many theological issues like divine revelation. Warfield’s theological engagement focused especially on theological issues like creation, 2

Peter Heslam has suggested that these differences lie primarily in how comprehensively the Enlightenment had penetrated Continental Europe, a claim that Bartholomew agrees with. See Peter S. Heslam, “Architects of Evangelical Intellectual Thought: Abraham Kuyper and Benjamin Warfield,” Themelios 24, no. 2 (1999): 3-20. See also George Marsden, Understanding Fundamentalism and Evangelicalism (Grand Rapids, MI: Eerdmans, 1990), chap. 5.

6

Geoffrey H. Fulkerson and Joel Thomas Chopp

providence, and supernaturalism, on the one hand, and the unity and nature of the human race, on the other. Joshua Jipp critically examines Rudolf Bultmann’s demythologization project and how developments in the natural sciences both motivated and influenced his interpretation of Scripture. For Bultmann, myth and science are not concerned with the same tasks or concerns. Whereas mythical thinking is guided by the question of one’s own human existence in its statements about the cosmos, scientific thinking speaks of the world rationally and from a distance. As Jipp shows, Bultmann held that Christians are not called to repristinate the mythical world picture of the biblical authors in part because modern science has made such a project impossible. Rather, Christians should understand Scripture is not giving us an objective picture of the world or making ontological truth claims but instead is expressing “faith in man’s present determination by God.” Ultimately, however, Jipp concludes that Bultmann’s project depends on a set of unhelpful and unnecessary bifurcations. Moving from Bultmann to Barth, we turn to another heir of the liberal Protestant theological tradition—one who also broke with it, albeit in very different ways. Nevertheless, the inclusion of Karl Barth in a volume such as this is something of a paradox. On the one hand, Barth is indisputably the most influential theologian of the twentieth century. Every figure in the final five chapters of this collection bears discernible imprints of his thought. On the other hand, Barth had little to say about the relation between theology and the natural sciences in the abstract, and even less about the relationship between particular scientific theories or developments and the Christian doctrines on which they might impinge. Katherine Sonderegger’s chapter sheds light on this curious tension. By investigating Barth’s category of “pure saga,” Sonderegger charts the similarities and differences between Barth’s doctrine of creation and the influential account laid out in Schleiermacher’s The Christian Faith. In contrast to Schleiermacher, Barth’s avoidance of either conversation or quarrel with the sciences is not because he construes theology as belonging to an entirely separate domain of inquiry. Barth’s doctrine of creation lays claim to the whole of reality, including the world that scientists inhabit and study. The relation between God and creature is a

Introduction

7

history, but not ordinary history in the everyday sense. Sonderegger’s chapter concludes by sketching how science, properly understood, might be a fruitful partner within Christian theology. T. F. Torrance, recognized as the greatest British theologian of his generation, is the first of several theologians downstream of Barth. In Kevin Vanhoozer’s chapter, Torrance appears as a “kataphysical” theologian— that is, a thinker in search of the “deep, primordial harmony” between being and knowing, between the object and structure of reality and our way of apprehending it. On this basis, Torrance finds agreement among strange companions: patristic thinkers like Athanasius and Philoponus, and modern physicists like James Clerk Maxwell and Albert Einstein. The Trinity, specifically the christological teaching of homoousios, is Torrance’s key not only for theology but for understanding all of reality. It both denies any separation of the way things appear from the way they are and resists any reduction of reality to mere appearances. While this chapter gives extensive attention to what is commonly called epistemology (the study of knowledge, scientia), it will also become clear that much more is at stake than method. In the question of knowledge, for Torrance and Vanhoozer alike, we are confronted with the question of discipleship, which is to say our way of life in the world. Jürgen Moltmann’s theology is one of the more revisionist of the theologians represented in this volume, especially related to his doctrine of God. The two primary foci of Williams’s essay are, on the one hand, Moltmann’s doctrine of God, which presents creation as an “open system,” and, on the other hand, what we might call the common environment of science and theology—which Williams is much more optimistic about. Rather than getting bogged down in the details and particular topics of this or that scientific theory, Moltmann demonstrates a form of theology that engages the broad patterns, wider picture, and deep meaning of nature. Moreover, the unity of science and theology is “forged in wisdom,” as they collectively attend to their shared history of this one world. Perhaps more than any other theologian surveyed in this volume, Wolfhart Pannenberg embodies the spirit of engagement that inspires this book. According to Christoph Schwöbel, his work “presents the most sustained attempt by any systematic theologian of the twentieth and

8

Geoffrey H. Fulkerson and Joel Thomas Chopp

twenty-first century so far to integrate the engagement with the natural sciences in his overall theological enterprise and to show the creativity with which theology can relate . . . to the theory formation in the sciences.” Schwöbel situates Pannenberg in the midst of the science-theology discussions in Göttingen. Pannenberg’s primary scientific interlocutors are physicists: Newton and Descartes serve as critical foils, and Faraday and his idea of a “field of force” is a constructive theological alternative. His motivating concern is, on the one hand, an account of the natural world that has no room for God and theology (i.e., mechanistic accounts of nature, principle of inertia) and, on the other hand, an account of theology that operates independently of science and nature (e.g., gaps, mere revelation, experience). Against both, Pannenberg sets forth a theology of nature, the exposition of which is the focus of Schwöbel’s chapter. Stephen John Wright’s chapter provides a careful exposition of Robert Jenson’s engagement with a surprising range of scientific theories. Wright shows how the characteristics that mark Jenson’s theology—the central role of the categories of narrative and history, his analogical treatment of time and eternity, his Christological orientation—provided fertile ground for engaging the sciences. Because both science and theology are dealing with the same reality, the world of creation, Jenson refused to construe the relationship between the two as either intrinsically conflictual or merely complementary. Neither is safe from the claims of the other. But with this risk, Jenson’s theology also shows that the two can assist each other with resources that would otherwise be unavailable. “Science can agitate theology into productive work,” Wright notes, but neither should theology attempt to merely limit itself to the constraints of the contemporary scientific narrative. Wright sketches three vignettes of Jenson’s employment of scientific theories for theological ends: Newton’s description of motion, Stephen Hawking’s distinction between imaginary time and real time, and Darwinian accounts of evolutionary adaptation. In each, Wright deftly uncovers the generative potential of Jenson’s rejection, on theological grounds, of construing science and theology as belonging to discrete epistemic realms. It is perhaps not surprising that Gunton, the youngest theologian of the volume, exhibits a theology shaped by many of the theologians who

Introduction

9

precede him, including Barth, Torrance, Pannenberg, and, on a personal level, Jenson (for a time, Jenson was Gunton’s doctoral adviser). In Murray Rae’s presentation of Gunton’s doctrine of creation, several common threads also reappear: creation as an article of faith, the nature of knowledge, the primacy of creation ex nihilo, the intellectual implications of theology’s affirmation of contingency. Rae especially sounds a trinitarian note, and he uniquely situates science as a human cultural enterprise. Rae, following Gunton, is as much interested in science as a human endeavor as he is in any particular body of knowledge. Prominent Themes and Threads Despite the disparate range of theological approaches and confessional commitments in the figures examined in this volume, several recurring themes can be discerned among them. Perhaps one of the more surprising commonalities is the extent to which they looked to theologians throughout earlier periods of the Christian tradition for resources to deal with distinctly modern problems raised by the natural sciences. B. B. Warfield drew on the Reformed scholastic Johannes Wollebius and his distinction between creation and mediate creation in his evaluation of evolutionary theory. Torrance appealed to Athanasius for his understanding of the relation between science and theology; both Torrance and Gunton find in the sixth-century theologian and physicist John Philoponus a compelling illustration of the generative potential of allowing theological commitments to shape scientific theorizing. Pannenberg invokes Duns Scotus’s account of contingency as a corrective to the mechanical philosophies of the eighteenth and nineteenth centuries. The majority of the figures examined here could not be charged with what C. S. Lewis called “chronological snobbery.” It appears that theologians of the past still have new things to teach us—even within the particular sets of questions and concerns that animate modern, scientifically engaged inquiry. The range of scientific theories and points of contact with theology are also broader than those concerned with human and universal origins. The theological implications of Isaac Newton’s physics make appearances in chapters six through ten. Newton’s, Einstein’s, and Hawking’s theories of time are explored and put to theological use by Torrance, Pannenberg,

10

Geoffrey H. Fulkerson and Joel Thomas Chopp

and especially Jenson. Faraday’s concept of field of force, Descartes’s formulation of the principle of inertia, and James Clerk Maxwell’s discovery of the speed of light all make appearances. Questions related to evolution and human origins are present as well—Kuyper on Herbert Spencer’s and Ernst Haeckel’s social deployment of the theory, Warfield’s complex engagement with the theories of his time, Barth on the vexed question of the “historicity” of Adam, Jenson on evolution and prehuman ancestry. Why, we might wonder, did one theologian engage Newton and another Darwin? And how should we who are not physicists discern between Newton and Faraday, or we who are not biologists between Darwin and Spencer? Moreover, what imaginative judgment allowed Torrance to connect relativity with Nicene Christology—and is it warranted, either for theology or for science? Can we find any common theological judgments that exercised these theological moves, or are their motivations for various ressourcement theologies as resistant to ordered reflection and instruction as the range of topics that they addressed? Several of the chapters also attend to the philosophical assumptions and commitments latent in modern scientific theorizing. Enlightenment figures that shaped scientific concepts such as Descartes, Leibniz, and Kant are present. The shift away from a broadly Aristotelian metaphysical framework is felt as well. In at least some cases, attention to the philosophical issues involved revealed that the apparent conflict between science and theology was grounded in prior competing (often, unexamined) metaphysical commitments. The influence of the mid-twentiethcentury Personalists is also increasingly evident downstream of Barth and Michael Polanyi, with resonances in the emphasis on story, history, ethics, and cultural criticism in the final five chapters. These philosophical interlocutors also bring us to one of the most prevalent threads throughout the chapters. In their engagement with God’s creation, both scientists and theologians share a commitment to “reality.” If science and theology both deal with one world, then genuine concord should be both possible and expected, but also genuine conflict.3 With some notable exceptions, the majority of the theologians Although discussions about “concordism” do not explicitly appear in these chapters, all the authors (Bultmann perhaps being the lone exception) hold to a kind of soft “concordism.” This is a natural consequence of the fact that they believe science and theology share a common world.

3

Introduction

11

presented here model what Peter Harrison calls “soft irenicism,”4 which emphasizes the historical contingency of both concord and conflict between science and religion. Harrison writes, Peace is a good thing, but it occurs because at that particular time the relevant science just happens to not conflict with religion. It might be the case, for example, that evolutionary theory does not conflict with a Christian view of creation. But for advocates of the soft irenic position, this position is not derived from any overarching principle about the necessary relations between science and religion, and no generalization about science-religion relations will follow from it. It is just that examination of the relevant scientific and religious doctrines yields, in this particular case, no evidence of conflict.5

Following the soft irenic position that Harrison identifies, potential concord or potential conflict between science and religion must be assessed on a case-by-case basis, based on the material claims themselves.6 These chapters provide just such examinations, with an emphasis on the historically contingent reception of scientific theories and equally contingent and imaginative theological responses among a handful of influential Protestant theologians.

Peter Harrison, “Is Science-Religion Conflict Always a Bad Thing? Augustinian Reflections on Christianity and Evolution,” in Evolution and the Fall, ed. William T. Cavanaugh and James K. A. Smith (Grand Rapids, MI: Eerdmans, 2017), 206. Hard irenicism operates with the assumption that conflict between science and religion is not possible in principle—whether because the two deal with entirely independent spheres or because “truth cannot contradict truth,” and thus if we can be reasonably sure about the deliverances of one discipline, the other discipline must modify its claims. Harrison notes that in practice, hard irenicism often results in a one-way direction of traffic, modifying theological beliefs to fit prevailing scientific theories. 5 Harrison, “Is Science-Religion Conflict Always a Bad Thing?,” 207, emphasis added. 6 “This stance also prompts us to look closely at the details of various scientific claims, with the possibility that some aspects of the general theory might be acceptable, but others not. Specifically, in the case of evolutionary theory the argument would not be that there is a scientific consensus about the truth of evolution and that therefore Christian thinking must adapt itself to this reality. Rather, it would be a matter of scrutinizing every element of the theory, its variant forms and their implications, and considering whether all or some or none were compatible with core Christian beliefs.” Harrison, “Is Science-Religion Conflict Always a Bad Thing?,” 208. 4

1 W I L L IA M BU RT P O P E ( 1 8 2 2 – 19 0 3 ) Primary and Secondary Creation Fred Sanders William Burt Pope was born to devout Methodists who had recently emigrated to Nova Scotia. The family returned to England in 1826, and in 1840 William entered the Wesleyan Theological Institute at Hoxton. After twenty-five years of active ministry, in 1867 Pope took up the position of tutor in systematic theology at Didsbury College, Manchester, which he retained until his retirement in 1886. Pope’s three-volume Compendium of Christian Theology remains one of the most comprehensive articulations of Christian doctrine in the Wesleyan tradition.

W

illiam Burt Pope was rightly described by a mid-twentieth-century commentator as “pre-eminently the Methodist theologian of the nineteenth century.”1 In 1985 Alan P. F. Sell added that “even today the last four words could be omitted without injustice to anyone else”; that is, Pope continues to stand as the preeminent theologian in the entire Methodist tradition. Sell goes on to characterize Pope as “the warmly devotional exegete, who brings from the store of scripture things new and old, and builds them into an impressive system.”2 Indeed, Pope was not only the greatest doctrinal theologian ever to take up the task of teaching Christian theology from the point of view of the Wesleyan revival movement’s spiritual core, but he ranks among the finest practitioners of Christian theology from any confessional group in modern Charles J. Wright, “Theology and Theological Tutors at Didsbury During a Hundred Years,” in Didsbury College Centenary, 1842–1942, ed. W. Bardsley Brash and C. J. Wright (London: Epworth, 1942), 51. 2 Alan Sell, “An Englishman, an Irishman and a Scotsman . . . ,” Scottish Journal of Theology 38, no. 1 (February 1985): 41-83. 1

14

Fred Sanders

times. To explore his three-volume Compendium is to engage in serious, high-level theology with one of the master practitioners.3 Pope excelled at stating the large, main ideas of the Christian faith, which makes it difficult to summarize what is distinctive about his theology. Thomas Langford hazarded this characterization: “The central idea in Pope’s thought was that of divine grace as effected in human life by the Holy Spirit.”4 If that just sounds like “mere Christianity” rather than an identifiable school of thought, Pope would not be upset by that characterization. He was trying to fill the office of a theological teacher, passing on what the church has always taught, and he wanted his work to be judged with questions like, Is it clear and memorable? Does it lead readers into a deeper understanding of Scripture? Is it a faithful restatement of the great tradition of Christian thought? Does it spend the right amount of time or the right number of pages on the most important things, putting less important items in subordinate places? Starting on the first page of his Compendium, Pope admonishes his readers of the dignity and sanctity of the study of theology: “It is A DEO, DE DEO, IN DEUM: from God in its origin, concerning God in its substance, and it leads to God in all its issues.” With a meaning far more than just etymological, Pope declares of theology, “His NAME is in it.”5 Hence every branch of this science is sacred. It is a temple which is filled with the presence of God. From its hidden sanctuary, into which no high priest taken from among men can enter, issues a light which leaves no part dark save where it is dark with excess of glory. Therefore all fit students are worshippers as well as students. . . . The remembrance of this must exert its influence upon our spirit and temper in all our studies. Who shall ascend into the hill of the Lord? or who shall stand in His holy place? He that hath clean hands, and a pure heart.6

William Burt Pope, A Compendium of Christian Theology, Being Analytical Outlines of a Course of Theological Study, Biblical, Dogmatic, Historical. 3 vols. The first volume of the first edition appeared in 1879; I will be citing the second edition, revised and enlarged (New York: Phillips and Hunt, 1881). 4 Thomas A. Langford, Practical Divinity: Theology in the Wesleyan Tradition (Nashville, TN: Abingdon Press, 1983), 68. 5 Pope, Compendium, 1:4. 6 Pope, Compendium, 1:5. 3

William Burt Pope (1822–1903)

15

The tone is obviously homiletic, but the work throughout is conceptually rigorous, which delivers it from being naively devotional. It does not degenerate into a weakly expressed series of merely edifying thoughts. Rather, those who have immersed themselves at length in the Compendium can testify that Pope’s integration of theology and spirituality is more nearly patristic than pietistic. Pope breathes the exalted air of the great theological tradition. Langford says there was very little in Pope’s theology that could not already be found in Wesley or the Methodist theologians Clarke and Watson. “The distinctive quality of Pope’s writing lay in his style of expression, his lucidity, and his completeness.”7 Alan Sell adds that Pope “was ever the constructive systematiser”8 who found a place for everything that had been scattered here and there in the Methodist tradition, and he gave great attention to the grand lines of the doctrinal system. Why, if Pope is such a virtuoso of the grand tradition of Christian theology, has his fame declined so precipitously? Why is he neglected and left unread? It is tempting to say that he was too good a theologian for the time he was born in, and perhaps too good for the people he sought to serve. He produced a serious Methodist Christian theology at a point in history when the Methodist movement was tending in some quarters toward a less serious, and in all quarters toward a less doctrinal, profile. He also produced an irenic synthesis marked by balance and proportion at a time when there was felt need for short, sharp polemics on the frontier with Calvinism. Pope was calmly and coolly anti-Calvinist, eager to show that the main lines of the Christian tradition were on the side of the Methodists (or, to put it the other way around, that the Methodist movement was well aligned with the great Christian tradition). He took the long view and wrote with the goal of forming the overall theological understanding of students. Even his own students, though, were unlikely to reach for their class notes when preparing for a theological conflict. Langford hypothesizes that one of the reasons for the eclipse of Pope’s work is that Pope found his voice in the mid-nineteenth century but Langford, Practical Divinity, 69. Sell, “An Englishman,” 58.

7 8

16

Fred Sanders

published later in the century. Things had changed already in his lifetime: Darwinism had emerged in that time, and many theologians became convinced that Christianity had to be restated within a thoroughly evolutionary framework. The tide of historical criticism of the Bible was rising rapidly during these decades, and again the younger generation was certain that responsible theology had to take the assured results of the latest critics into account. Pope, for whatever reasons, serenely declined the invitation to take part in the panics and scrambles that animated so much Victorian intellectual life. Langford points out that in his vocation as a teacher, Pope focused on communicating what he had already learned, but he was “isolated from the newer currents in British intellectual life—Darwinism and idealistic philosophy.”9 On this view, Pope was not engaged in the swirl of Victorian struggles with religious doubt, the new dynamic of biblical criticism, the changing philosophical scene with the rise of idealism, or the transforming power of evolutionary ideas. His position was formed prior to the 1860s, the critical period for many of the issues, and although the Compendium was published in 1875–76, he was not responsive to these new currents. Catholic in his range of sensitivity to traditional Christian positions, Pope was uncongenial toward contemporary developments, although a slight familiarity with Friedrich Schleiermacher’s theology is evident.10

It may well be that Pope was simply out of touch with the recent developments. But it may also be that while he was adequately informed about them and had taken their measure, he had decided that they simply did not force themselves on the consideration of a theologian in the same way that the large, central doctrines did: Trinity, incarnation, atonement. Langford acknowledges that Pope’s lack of engagement with contemporary controversies was his “conservative disposition and his

Langford, Practical Divinity, 69. Langford, Practical Divinity, 69-70. For further remarks on Pope and Darwinism, see Thomas A. Noble, “Darwin and Theology,” Didache: Faithful Teaching 15, no. 1 (Summer 2015): 1-16. David N. Livingstone also provides a brief comparison between the attitudes of Pope and B. B. Warfield: Pope “similarly adopted the theory of evolution, while holding that it had not provided an explanation of the emergence of life or the emergence of intelligence.” David N. Livingstone, Darwin’s Forgotten Defenders (Grand Rapids, MI: Eerdmans, 1987), 135.

9

10

William Burt Pope (1822–1903)

17

tight focus on biblical truth.”11 Other commentators have assumed that Pope was “keenly aware of the contemporary challenges to the Christian faith” and even “often concerned with these apologetic themes.” Nevertheless, his impact was blunted and “his stature as an apologist . . . reduced by his tendency to brush aside extreme atheist arguments without much consideration. When the materialists and positivists say that God is imaginary, this is too much for Pope, who ‘has nothing to say to such impiety and skepticism.’”12 Of course, to say aloud that one has nothing to say to impiety is itself to say something to impiety by adopting a definite rhetorical strategy. Principled disengagement is still disengagement. As a British Methodist, Pope’s church connection was numbered among the dissenting bodies. Perhaps it was in compensation for this that his overriding concern as a leader was to keep the Methodist theological tradition solidly connected to, and in conspicuous continuity with, the main stock of historic, traditional Christian doctrine. For various reasons, some Wesleyans and Arminians have interpreted their origins in the great revivals as warrant for interpreting their distinctive theology as the documentation of something brand new that God is doing in the world; at its worst, the tradition seems to presuppose that theology needs to start more or less from scratch now that the Methodists have arrived. Sometimes that attitude is expressed simply as antiCalvinism, reacting against the five-point synthesis of Dordt and its attendant emphases. But under certain conditions, the Wesleyan theological impulse can interpret Calvinism as the nearest edge of a vast, Augustinian complex of errors. What follows from this is usually a kind of intellectual sectarianism, as Methodist theology sets itself up as a protest movement against Augustinianism. Down this road lies an antipathy to the theology of the great central tradition of Western Christian thought. We need not accept Augustine uncritically in order to acknowledge his importance, and it is easy to see how a commitment to working around his theological legacy could become an impulse toward Langford, Practical Divinity, 69. William Strawson, “Methodist Theology, 1850–1950,” in A History of the Methodist Church in Great Britain, ed. Rupert E. Davies, A. Raymond George, and Gordon Rupp (London: Epworth Press, 1983), 3:185. Strawson quotes Pope from Compendium, 1:60.

11 12

18

Fred Sanders

marginalization. This impulse has not been good for Wesleyan theology. Pope’s Compendium is especially designed to show how Methodist theology is one of the voices that sounds out alongside the others in the same Christian family of doctrine. He is sometimes harder to read than Watson and Miley, two other accomplished and nonsectarian theologians of the Wesleyan tradition. Pope writes more artfully, with greater allusiveness. For constructive, doctrinal purposes, Pope’s commitment is to the grand vision of the unity and continuity of Christian doctrine. Pope was once commissioned to write an essay explaining the distinctiveness and peculiarity of Methodist doctrine, a task he was glad to undertake but that he necessarily shaped to his own ends. “We have to say a few words upon certain peculiarities in the doctrinal position of Methodism,” he begins. But he immediately changes the subject: “But it is a pleasant preface to dwell for a moment on the broad expanse of catholic evangelical truth, concerning which it has no peculiarities, or no peculiarities that affect Christian doctrine.”13 In his next sentence he launches into a summary of the shared creation theology of the Christian church: “To begin where all things have their beginning, with the being, triune essence, and attributes of God; his relation to the universe as its Creator and providential Governor; his revelation of himself in nature: this supreme truth it holds against all atheism, antitheism, pantheism, and materialism.”14 It is here, with the shared Christian doctrine of creation and its antithesis to other cosmological theories, that we turn from introducing our neglected theologian to taking up his most significant proposal for a Christian doctrine of creation. Primary and Secondary Creation in Pope’s Theology One of the characteristic features of Pope’s doctrine of creation is a programmatic distinction he offers between the divine act of primary creation (God calling all things into existence) and the divine work of secondary creation (the formation of an ordered universe). The former is William Burt Pope, “Methodist Doctrine,” in The Wesley Memorial Volume, ed. J. O. A. Clark (New York: Phillips & Hunt, 1881), 173. 14 Pope, “Methodist Doctrine,” 173. 13

William Burt Pope (1822–1903)

19

the realm of metaphysical inquiry and apologetic argumentation; Scripture chooses to say little about it, and science can say nothing in principle. The latter, secondary creation, is the domain of both the biblical account and of scientific investigation, and it also stretches forward into the doctrine of providence. The distinction between primary and secondary creation had been broadly shared among the theologians of Protestant orthodoxy in the centuries before Pope and has roots in medieval and patristic thought. Characteristically, Pope was not attempting originality in his use of the distinction, but he did carry it through his entire account of creation with special consistency. Early in his discussion, Pope says of the doctrine of creation that “the revelations of Scripture on this subject may be distributed under the two heads of the Creator in regard to the act of creation and the several orders of the creatures as the result of His creating act.”15 What Pope emphasizes even here is the singleness of God’s one creative act, on the one hand, and the multiplicity of ordered, layered, sequenced creatures that result, on the other. That is, he contrasts the unity of the creative act with the multiplicity of created entities. Both are true and posited by Scripture. But they demand distinct treatment, and it is this distinct treatment that orders Pope’s account of creation. With reference to the doctrine of God, primary and secondary creation display the leading divine attributes differently. “The creating act displays the glory of [all] the Divine attributes, but freely as an act of will, and with the diffusion of happiness as one end attained by the resources of infinite wisdom.”16 God does not create out of necessity, but when he chooses to create, the act he freely carries out displays his glory. The sheer act, however, or “absolute creation,” shows itself as “the effect of omnipotence”; it shows primarily the sheer power of God to do this great thing, and “the origination of creaturely existence is a mystery which is revealed for adoration only, no other account being given or possible but the all-sufficiency of the Creator.”17 Absolute creation makes divine power known but does not, considered by itself, communicate much about the Pope, Compendium, 1:361. Pope, Compendium, 1:362. 17 Pope, Compendium, 1:362. 15 16

20

Fred Sanders

divine wisdom or love. On the other hand, “Secondary creation, or the formation of the material part of the universe into order, exhibits Divine wisdom also and love as preparing the scene of Providence.”18 Primary creation is, of course, not unwise or unloving; it is simply too vast and unformed a notion to be an effective display of anything but omnipotence. It is the detailed, layered, sequenced, and interrelated character of secondary creation that manifests the wisdom of God, setting up the ordered, sequential dynamics in which his love is made known. Pope is concerned to be as clear as possible about how the divine attributes are manifested in creation because, in his view, everything about the discipline of theology must ultimately be traceable to its foundation, which is “the meditation and study of the Divine Attributes.” Theology is “by the very term [that is, as logos about theos] the doctrine of God contemplated in Himself and in His universal relations, or in the universal relation of all things to Him.”19 The distinction between primary and secondary creation gives him the opportunity to draw attention to the overall structure of the divine attributes: It is enough to say here that the omnipotence of God, as the outward manifestation of His interior all-sufficiency, is enough for the original production of matter in what may be called absolute creation; that His wisdom and power are seen in the secondary creation or formation of matter into worlds; and that the end of all is the expression of the Divine perfections or their reflection in the works of His hands.20

Characteristically, Pope is working with several levels of distinctions in place but is not rehearsing them at each point. Divine all-sufficiency is an immanent attribute of God; manifested outwardly it is omnipotence. This immanent all-sufficiency that is manifested outwardly as omnipotence suffices for primary creation; or, tracing the steps back the other way, the sheer fact that God creates anything at all is a work of omnipotence grounded in divine all-sufficiency. Pope the Methodist is concerned enough about Reformed theological distortions that he exercises some caution in the account he gives of Pope, Compendium, 1:362. Pope, Compendium, 1:353. 20 Pope, Compendium, 1:364. 18 19

William Burt Pope (1822–1903)

21

omnipotence and sovereignty. He may have a strategic maneuver in mind when he seeks to do justice to the theme of divine power in this section on “creation proper.”21 In Pope’s view, “it is only in the Divine All-sufficiency that we can find the ground of the origin of all things that exist not being God Himself. In this we must be content to seek the possibility of all forms of being.” In other words, since it is incumbent on all Christian theologians to confess God’s omnipotence, Pope takes the opportunity to make his strongest statements about omnipotence while he is describing primary creation. This preemptive move would not be enough to fully satisfy Reformed critics, of course; nor does it relieve Pope altogether of the requirement to confess divine sovereignty in the details of secondary creation in the world process. But it does relieve some of the stress placed on the doctrine of omnipotence by Reformed accounts by giving Pope room to confess omnipotence in a sheer, undetermined, absolute way: “With God all things are possible [Mt 19:26]: that is, all things possible may, at His will, become actual.”22 For creation proper, Pope builds both a positive and a negative case. The positive case is directly constructed from the affirmations of Scripture. Pope argues exegetically from predictable texts like Genesis 1 (“In the beginning God created”), John 1 (“Through him all things were made”), and Colossians 1 (“In him all things were created: . . . He is before all things, and in him all things hold together”). But he also argues from Romans 1, which he calls “a very remarkable passage,” in which “St. Paul declares the possibilities of God to be ta aorata autou, the invisible things of Him.”23 In other words, “What He in the freedom of His omnipotence brings into visible existence proclaims His eternal Power and Godhead.” Pope examines the relation between power and Godhead carefully: “The dynamis here preceding, and measuring, and, as it were, determining the theiotes, while, on the other hand, the theiotes, or divineness of God, is the substratum of that dynamis, the resources of which are infinite.”24 What interests Pope here is the essential recess of divinity in itself as the “Creation Proper” is a heading at Compendium, 1:364. Pope, Compendium, 1:364. 23 Pope, Compendium, 1:365. 24 Pope, Compendium, 1:365. 21 22

22

Fred Sanders

fund of creative power. He may in part be guarding against an abstract notion of power overdetermining the account of God’s particular identity, but it is characteristic of Pope that his interest is mainly in spelling out the full meaning of the exact words of Scripture. Similarly, Pope finds much in the language of Hebrews 11: “By faith we understand that the universe was formed at God’s command, so that what is seen was not made out of what was visible.” Pope fastens on the phrase “the worlds were framed” as a statement about secondary creation—that is, world-framing. Likewise, Pope takes the phrase about the appearing of visible things from that which is not as a statement about primary creation, or the absolute origin as such. He summarizes, “The construction and the absolute origination of all things seen are, in fact, separated, and then united. The creative word of God is set over against both.” Pope finds in this passage the most careful statement in which Scripture approaches the formulation “ex nihilo,” and he believes that Scripture makes the distinction between world-framing, on the one hand, and the appearing of things, on the other hand, so that “faith may lay hold of the truth, which reason cannot penetrate, that the created universe did not spring by development from things previously existing, but from the invisible creating power of One afterwards referred to as . . . Him Who is invisible [Heb 11:27].”25 According to Hebrews (according to Pope), primary and secondary creation are distinguished and then united precisely in order that the Christian relation to creation will be one of faith, one that draws the conclusion that created phenomena have a depth in God which transcends all possible created relations. This is strikingly similar to what Robert Sokolowski has described as making the Christian distinction, the theological inference drawn from worldly phenomena that sets Christian revelation apart from paganisms of every kind by gesturing toward the true God.26 Pope’s final word on this is telling: “With this the revelation of Scripture has spoken its last word, after which the first word of science must begin.”27 Pope, Compendium, 1:366. Robert Sokolowski, God of Faith and Reason (Washington, DC: Catholic University of America Press, 1995), chap. 1. 27 Pope, Compendium, 1:366. 25 26

William Burt Pope (1822–1903)

23

Thus far we see the positive, exegetical case for creation proper, or primary creation. Pope also builds a negative case. His negative case consists of arguing that “Scripture precludes any other doctrine than that of an absolute creation of all things by the direct act of the Divine will.”28 In his view, there is simply is no other way to preserve “the infinite and to us unthinkable chasm between necessary being and existence phenomenal.” While he admits that “the Bible does not say, in philosophical language, that the Unconditioned One remains the Unconditioned while He created the conditioned, or that the Necessary Being cannot have other necessary existence, co-eternal with Himself, which He forms into the universe,”29 Pope nevertheless affirms that this is what Scripture means by what it says. He canvasses all the other options for a relation between God and the world and not only finds them unacceptable but finds their unacceptability to constitute the negative proof for the correctness of the Christian view. In this next section of the Compendium of Christian Theology, Pope takes up each alternative possibility in turn and explains the problems they entail. He devotes six pages to pantheism, eight to polytheism, four to dualism, and ten to materialist atheism. The root error of pantheism, according to Pope, is “identifying . . . God with the universe,” or more tellingly, “making God supreme in it without being its Creator.”30 Pope is eager to distinguish a range of types of pantheism (“All Pantheism is not the same Pantheism”31), distinguishing the religiously motivated Hindu systems from the speculative system of Spinoza. Of course, for the purposes of his argument, they all offer various forms of the same root error. Polytheism comes in for a long discussion, although Pope admits it is not an error against which Christian theology needs to contend much in the modern West. In addition to having some genuine interest in the histories and cultures that produced a range of polytheisms, Pope finds that “the study tends greatly to serve the cause of the Christian religion by showing the incomparable superiority of the records of revelation.”32 Pope, Compendium, 1:366. Pope, Compendium, 1:366. 30 Pope, Compendium, 1:367. 31 Pope, Compendium, 1:367. 32 Pope, Compendium, 1:381. 28 29

24

Fred Sanders

Rejecting polytheism is so easy for Pope that this section’s main value is its exploration of varieties of religious mythology and what that variety reveals about human nature. Pope considers materialism the polar opposite of pantheism, “the philosophical or scientific antagonist of the Scriptural doctrine of the Creator and creation.”33 Materialism’s error is that it “gives matter the pre-eminence, as the only substance that is, and regards what men call God as the unknown law by which that substance is governed in all its evolutions.”34 When he discusses the origin of human life, Pope laments that any materialist theory makes man an organism in which matter exhibits its perfection in the phenomena of thought and conscious personality. Modern speculations on this subject differ generally from the ancient, in consequence of their being constructed on a theory that does not necessarily exclude a personal God, the origin of all life. Placing Him at the ultimate point where life originated, they regard the evolution of all the forms of life as the operation of forces impressed upon matter, or constituting matter itself.35

At those points where evolution depends on materialist metaphysics, Pope unleashes his full scorn on it. Alan Sell says that “the impersonality of the evolutionist’s fundamental principle was what especially struck Pope,”36 and it is this that draws out his fiercest denunciations. In one address, Pope confronts the materialist philosophy of evolution thus: Men persuade themselves to accept a law of silent, ceaseless evolution ruling in the economy of things, to which “one day is as a thousand years, and a thousand years as one day.” See how patiently they wait upon the slow travail of millenniums and cycles of ages; watching the disappointments of nature as feebler types perish, and allowing vast periods of time for every new and better feature to be stamped on the ascending creature. Their language is “thou law”—not “Thou Lord”—“in the beginning didst lay the foundation of the earth and the heavens are the work of thine hands.” I say, let us learn to confirm our faith by their irreverent unbelief. Pope, Compendium, 1:385. Pope, Compendium, 1:385. 35 Pope, Compendium, 1:431. 36 Sell, “An Englishman,” 46. 33 34

William Burt Pope (1822–1903)

25

While they abase their minds before a dread irrational necessity or force, and patiently wait upon it, let us humble our minds before the eternal majesty of wisdom in “the patience of the saints.”37

With his rejection of alternative systems, especially materialism, Pope concludes his treatment of absolute creation, or primary creation. It is a treatment in which Pope strategically makes the most of the doctrine’s vastness and lack of complexity. Although the doctrine of primary creation is characterized by a certain featurelessness and lack of detail, Pope makes a virtue of that by gathering together all of his broadest, most universal statements about creation. This provides a background against which he can carry out his treatment of secondary creation. Secondary Creation: World-Forming When Pope turns in his exposition to secondary creation, he takes up the biblical teaching about the formation of the varied contents of the world. “It is necessary to establish a distinction between the first production of matter and its subsequent elaboration, if such a term may be used, into the Cosmos, which brings us into the region of Cosmogony or Cosmology.”38 Where he refers to the product of primary creation as being, or that which is, he consistently uses the word world to mean a formed or crafted set of creaturely things in relation to each other and to God. Though in other contexts he might use language about creating the world in a looser sense that could refer to primary or secondary creation, when he has the distinction in mind he is careful to employ the word world only for the latter. Pope thinks this distinction is intimated in Genesis 1’s use of “the double expression Created and Made,” which he says “seems significantly to indicate a distinction which is not clearly defined.”39 The distinction would be that God creates being but makes the world. Pope has much to say in interpretation of the Mosaic cosmogenesis, the story of creation in Genesis 1–2. His exegesis of the passage is rich William Burt Pope, Sermons, Addresses, and Charges (London: Wesleyan Conference Office, 1878), 79. Cited by Sell, “An Englishman,” 46-47. 38 Pope, Compendium, 1:396. 39 Pope, Compendium, 1:396. 37

26

Fred Sanders

and interesting, and he goes out of his way to consider the possibility that this is only the account of this world of ours because he takes the Bible’s phraseology to at least permit, but perhaps even to suggest, multiple worlds. These could be either further back in time and buried by cycles and sequences of development, or so remote in outer space that they need not be considered part of our world, regionally speaking. He also has much to say about angels, whose spiritual existence populates the vaster cosmology within which he situates the earthly account of creation. Pope’s willingness to limit world to a region nearby is fascinating because it seems to leave most of the universe out of his purview. This restriction could position Pope’s cosmology and Bible interpretation as a more modest project, articulating no commitments about outer space, alien life, and so on. We skip over these possibilities in our exposition to make some observations about his view of humanity, because that is where the primary/secondary distinction exerts itself again. Throughout his exegetical investigations in Genesis, Pope adopts an intentionally defensive stance against modern biological and anthropological accounts. Because “speculations as to the Origin of Man upon the earth have been more or less bound up with those on the origination of life generally,”40 Pope approaches the subject of human evolution by placing it in a broad context, which entails that his irritation with materialism resurfaces here and finds its focus in the question of the human soul. Materialistic evolutionary accounts for “all, whether they intend it or not, practically denying the creation of the human soul or spirit as a substance distinct from matter.”41 In his view, It is impossible so to state the theory of evolution as to preserve the integrity of the higher element in man’s nature. But the true theory of that nature requires that something was superadded to the physical and immaterial life that lay behind it in the history of the creation. The Scriptural account is plain and express: man was created in the image of God . . . modern science will never find rest until it is acknowledged.42 Pope, Compendium, 1:431. Pope, Compendium, 1:431. 42 Pope, Compendium, 1:432. 40 41

William Burt Pope (1822–1903)

27

The echo of Augustine’s Confessions here is probably intentional: because God made us for himself, science is restless until it finds its rest in recognizing God’s image in an immaterial soul. In his theological anthropology, Pope is confident that the Bible has clearly revealed the nature of humanity and its origin. As such, the theologian’s responsibility to clear revelation is straightforward. “This Divine account of man’s origin displaces every other devised by man’s science.” Therefore, “accepting the testimony, as we believe it, of the Creator Himself, we have only to stand on the defensive.” Pope would probably have recognized an intellectual responsibility to reexamine his interpretation of Genesis if some other line of evidence forced itself on his rational powers, but he emphatically denied that any other line of evidence had yet done so: “It may safely be said that no other hypothesis of the production of mankind has yet proved its case.”43 Here we begin to see Pope’s attitude toward the sciences and his position on faith and reason more broadly. Alan Sell’s historical work on Pope offers some good comparative insights on this front: “That the nineteenth century was a time of theological reappraisal is well known. The question of the starting point of theological enquiry; the challenge from evolutionary thought and biblical criticism; matters historiographical and ecclesiological”44 all come to the fore in the intellectual climate of the time. Pope’s outlook is congenial to rational, philosophical argument, but philosophical may be the key word, because Pope seems more responsive to arguments from actual philosophy than to assertions from science. He accepts, for example, the utility of theistic arguments: “The Being of a God is at once an innate idea and a truth demonstrable and to be demonstrated.”45 Even here he admits that “there is a limit to their demonstrative force as human evidences: they require the enforcement of the Holy Spirit’s influence as Divine credentials.”46 According to Sell, Pope’s apologetic “stands ultimately for the priority of revelation. This, Pope, Compendium, 1:430. Sell, “An Englishman,” 41. See also A. P. F. Sell, Theology in Turmoil (Grand Rapids, MI: Baker Books, 1986). 45 Pope, Compendium, 1:234. 46 Pope, Compendium, 1:236. 43 44

28

Fred Sanders

after all, is the biblical stance.”47 Sell cites Pope’s statement from the early pages of the Compendium: “It is certain that it is more after the manner of the Bible to set out with the credentials of Revelation itself than to array a number of internal and presumptive evidences in its absence.”48 This may be the “manner of the Bible,” which Pope intends to follow broadly, but Pope also recognizes the difference between the Bible’s own words and the words of a Christian theologian who has to engage with the questions and concerns of later ages. He takes faith to be foundational, but he is no simple fideist. Sell attempts to locate him among more recent positions on the question of faith and reason: he considers Pope to be somewhere between evidentialism and presuppositionalism.49 Overall, while Pope employs theistic proofs, he seems to think that “the theistic arguments simply confirm what the believer has already perceived by the Spirit through the Word. In this respect the Arminian is a true son of Calvin.”50 On evolution, Pope accepts the idea of development broadly but “contends that the very intricacy of the processes, to which the scientists rightly point, requires belief not in impersonal force, but in the personal God.”51 Speaking before the British Association for the Advancement of Science, he said, “The revelation of natural science cannot contradict the revelation of spiritual science.”52 On subjects like this, there is a fideistic tone that creeps into Pope’s style, which can possibly be accounted for by Pope’s reaction against the infidelity spreading not so much in society at large but in the church and theological training centers of his time. Sell cites as an example of Pope’s concerns the way that “in a funeral address on the Reverend John Lomas he praised this theological teacher for impressing ‘upon a considerable number of students the claims of systematic Sell, “An Englishman,” 45. Pope, Compendium, 1:50. 49 Sell, “An Englishman,” 45. 50 Sell, “An Englishman,” 45. 51 Sell, “An Englishman,” 47. 52 Sell, “An Englishman,” 49. Sell quotes from W. B. Pope, “Jesus Anathema or Jesus Lord: A Discourse . . . on the Occasion of the Meeting of the British Association for the Advancement of Science,” (London: Wesleyan Conference Office, 1883) p. 21. This single sermon of thirty-one pages was printed and “published for the author at The Wesleyan Conference Office, 1883,” but Sell describes the copy he is consulting as being “bound within The Abiding Word.” 47 48

William Burt Pope (1822–1903)

29

or dogmatic divinity in opposition to the latitudinarian characterless negation of belief that has been creeping in among us.’”53 In his programmatic inaugural address for his post at Didsbury, Pope lamented that “wherever we turn our glance upon Christendom, we perceive the manifold signs of a steady, persistent, ruthless, and thorough determination to bring the Christian faith, and its holy documents, and its equally holy institutions, before the bar of a reason that will know nothing of faith.” The leading edge of this infidelity comes from teachers “seeking to rob the Old Testament Scriptures of the marks of their divinity as from God, and of their historical worth as from man.”54 This is the polemical context in which Pope perceived the need to emphasize, in theological training, a high doctrine of Scripture for both its intellectual and spiritual importance: “No man can be a genuine disciple of Christ who does not receive the Holy Oracles at His hands as a testimony to Himself given by His own Spirit to the prophets before He came, and by His own Spirit to the apostles after He departed.”55 It was the late nineteenth century, and Pope was fully aware of the need to stand against destructive currents of thought in theological education. Nevertheless, for all his insistence on the necessity of submission to the teaching of Scripture, Pope’s work is not obscurantist or tilted against input from the natural sciences. It is in no small measure Pope’s distinction between primary and secondary creation that gave him the ability to balance the claims of faith and reason in his account of creation. Primary creation is very much amenable to philosophical demonstration. We have seen that for Pope, theistic arguments show that there is a Creator, and he further believes that alternative views can be rationally argued to be incoherent or at least less satisfactory. On this side, the side of primary creation, Pope’s apologetic is nearly rationalistic; it is certainly evidentialist. Secondary creation, however, plunges us into the biblical Sell, “An Englishman,” 58-59. The funeral address he cites is in Pope, Sermons, Addresses, and Charges, 61. 54 Cited in Sell, “An Englishman,” 59. 55 W. B. Pope, The Person of Christ (London: Wesleyan Methodist Book Room, 1885), 42. In another context it would be worth distinguishing Pope’s bibliology from later varieties. This quotation, for instance, is from his book on Christology and forms part of his doctrine of Christ. On inspiration, he makes some careful distinctions about verbal inspiration at Compendium 1:189. 53

30

Fred Sanders

revelation about the making of this world of ours, and we must accept or reject the biblical testimony. Here our response to Moses is our response to God, the response of faith. On this side, the side of secondary creation, Pope’s tone is more nearly fideist. For Pope, the two approaches cohere precisely because the Christian doctrine of creation is composed of two parts, primary and secondary. If Pope’s approach to the doctrine of creation is successful, it gives him the benefit of rational proof of a creator, on the one hand, and a demand for faith regarding the genesis of the cosmos, on the other. In Conversation with W. B. Pope Even if a reader were to judge that Pope was not entirely successful in striking the right balance between biblical revelation and the natural sciences in his doctrine of creation, the distinction he draws between primary and secondary creation is one that could underwrite further attempts to do justice to the claims of both sides. And even if what Pope did in the middle and late-nineteenth century is no longer a direct model for our emulation in our own setting, his programmatic primary-­ secondary distinction could be pressed into service in our own context to meet challenges of our own time and enable a balanced Christian confession of creation. What follows are a few indications of the promise of constructive engagement with Pope’s doctrine of creation today. By way of preliminary consideration, there is the matter of warrants for making and deploying this distinction. To reclaim Pope’s primarysecondary distinction, we would need to clarify its appeal to Scripture. We should not expect to find metaphysical distinctions of this nature stated directly in the words of the Bible: What would the Hebrew and Greek idioms be for talking about unformed matter, or for the unconditioned, and so on? Pope himself uses the distinction as a conceptual tool for restating in very different terms what he takes Scripture to be presupposing in the way it actually speaks. More needs to be done to make more explicit the lines of thought in Scripture that generate the need for such a distinction. Without claiming that the Old Testament intends such a distinction when it alternates bara (create) and asah (make) in its creation accounts and elsewhere, the categories of primary and secondary

William Burt Pope (1822–1903)

31

may nevertheless serve well in teasing out the subtlety with which these words are woven together.56 Turning from Scripture to tradition, how much historical warrant does Pope’s distinction have? The distinction is present in a more Platonic idiom, deriving ultimately from the Timaeus and worked out in Augustine’s allegorical interpretation of “heaven and earth” as prime matter and pure form in Confessions 11. As this twofold account of creation made its home in medieval and Protestant orthodox theology, it was refined in various ways before it became the common property of numerous theological systems. What is unique in Pope is not the distinction itself but his programmatic use of it. The following is a necessarily brief list of some of the advantages to this programmatic use. First, the primary-secondary distinction is helpful for organizing the material of a doctrine of creation. As we saw in Pope’s own uses of it, this distinction buys the theologian adequate space for the variety of doctrinal concerns that need to be treated. John Webster has emphasized the importance of maintaining scope and balance in the Christian doctrine of creation, in which “disarray results from the hypertrophy or atrophy of a given element (as, for example, in theologies that reduce the doctrine of creation to teaching about created things, without adequate consideration of the Creator and his work).” Within this doctrinal field in particular, theologians must take care not to let one part “deform the whole, whose force depends in part upon the integrity of its constituents.”57 The difference between the Creator and the creation is absolutely fundamental for Christian confession, and it is a distinction that needs to be handled above all at the level of primary creation. As Robert Sokolowski has argued, any attempt to mark the absolute distinction between Creator and creature simply by pointing to particular created things will ultimately fail. Because “the Christian God is . . . so transcendent to the world that he could be, in undiminished goodness and greatness, even if the world were not,” he “can be distinguished from the world in this radical Bara in Gen 1:1, 21, 27, etc.; asah in Gen 1:5, 16, 25, etc.; both side by side in Gen 2:3 and 2:4. John Webster, “Love Is Also a Lover of Life: Creatio ex Nihilo and Creaturely Goodness,” in God Without Measure: Working Papers in Christian Theology, vol. 1, God and the Works of God (London: Bloomsbury, 2016), 100.

56 57

32

Fred Sanders

way.” Focusing attention on primary creation is a useful way of raising awareness of the radical character of this transcendence. It is not any part of creation that the Christian God transcends, but creation as such. “In contrast, the gods of pagan religion and the first principles of pagan philosophers are gods and principles for the world and they could not be without the world.”58 Second, the primary-secondary distinction affords some opportunities for clarity in handling the doctrine of divine attributes. Pope’s correlation of power with primary creation and wisdom with secondary is suggestive and promising. God’s sheer creative power is better grasped as the production of everything in general than as the crafting of all the particular things. Conversely, his creative wisdom is better seen in the intricate fit between means and end, part and whole, layers on layers and wheels within wheels, than it is in the mere fact that he made all things. The manifold wisdom of God shines out in the diversity of things made, and that is best explored under the heading of secondary creation, or world-crafting. Other divine attributes can also be illuminatingly assigned to one or the other of this broad distinction, not exclusively (as if God were not wise in primary creation or strong in secondary creation), but by way of instruction. The full range of attributes is necessary to do justice to biblical revelation: Ezekiel saw an exalted one seated on a throne high above the expanse, but he also saw the four multiform figures of the living creatures and their wheels. Third, the distinction opens up more space within the doctrine of creation that corresponds to giving more attention to the description of God’s own life in itself. It would be easy to skip over a discussion of primary creation, which is a more abstract side of the doctrine and has little immediate connection to the actual world in which we have our being. But this first stage of Pope’s two-stage doctrine of creation is a helpful intermediate step between the doctrine of God in himself and the doctrine of God toward us. Absolute creation provides an extra layer of consideration on the way up from the doctrine of creation to the doctrine of the creator. John Webster has shown how older schools of theology Sokolowski, God of Faith and Reason, x (preface to the new edition).

58

William Burt Pope (1822–1903)

33

were careful to keep several such distinctions in place within the doctrine of God; after describing a few of them, he asks rhetorically, “Why insert all these layers between God in himself and God’s temporal acts?” He answers that Christian theology does so “in order to characterize the agent of these acts, and so to come to understand and give due weight to both the acts themselves and their objects.”59 The task of describing primary creation plays an answering role on the creaturely side. Fourth, the doctrine of providence benefits from this distinction, which can do justice to God’s absolute and unconstrained power at the level of primary creation, without requiring divine omnicausality at the secondary level. Pope has a strong doctrine of providence, but he denies the appropriateness of talking about “absolute sovereignty” at work in history. Absolute means not related, but history is by definition about relation. It is the field of differentiated, particular creatures interacting with each other and their maker. The almighty, living God first exercises absolute power in primary creation, then wields his sovereignty in secondary creation by relating each particular thing to the others in series and networks of relation. Finally, Pope’s distinction establishes some helpful guidelines in the difficult theological field of divine action. We might say there are two broad perspectives about how God acts. On the one hand, the living God takes particular and definite actions within the course of history. On the other hand, we must confess that God is not simply one of the many agents operating within the world, competing with other agents for causal power. Both perspectives must be accounted for; both are true. Pope’s programmatic distinction could enable theology to recognize God’s transcendent agency at the level of primary creation and another, more narratable kind of agency within secondary creation. Recommended Reading Primary Works Pope, William Burt. A Compendium of Christian Theology, Being Analytical Outlines of a Course of Theological Study, Biblical, Dogmatic, Historical. 3 vols. 2nd ed. Revised and enlarged. New York: Phillips and Hunt, 1881. Webster, “On the Matter of Christian Theology,” in God Without Measure, 1:6.

59

34

Fred Sanders

———. A Higher Catechism of Theology. New York: Phillips and Hunt, 1884. ———. “Methodist Doctrine.” In The Wesley Memorial Volume, edited by J. O. A. Clark, 168-90. New York: Phillips & Hunt, 1881. ———. The Person of Christ: Dogmatic, Scriptural, Historical. 2nd ed. London: Wesleyan Conference Office, 1875.

Secondary Works Langford, Thomas A. Practical Divinity: Theology in the Wesleyan Tradition. Nashville: Abingdon Press, 1983. Sell, Alan P. F. “An Englishman, an Irishman and a Scotsman.” Scottish Journal of Theology 38, no. 1 (1985): 41-83. Shoaf, Robb Wicke. “The Theology of William Burt Pope: A Nineteenth Century Wesleyan Systematic.” PhD diss., Drew University, 1990.

2 A B R A HA M K U Y P E R ( 1 8 3 7 – 192 0 ) Enlightenment, Science, Worldview, and the Christian Mind C raig Bart h ol om ew Abraham Kuyper was one of the greatest Reformed thinkers and practitioners of the nineteenth and twentieth centuries. The son of a pastor, and a student at Leiden University, Kuyper was attracted to the liberal theology sweeping across the continent. However, while doing his doctorate in theology, he was converted through reading Charlotte Yonge’s Anglo-Catholic novel, The Heir of Redclyffe. Kuyper saw from the outset that Christianity related to all of life. In his first church he moved into the Calvinist camp but sought a form of Calvinism that was not ossified but rather supple and ­relevant to the great challenges of the day. Through the rest of his life he developed this neo-Calvinism in creative and comprehensive ways.

A

braham Kuyper’s major work as a theologian is his three-volume    Encyclopedia of Sacred Theology. This remains a fecund work for theology and science. In order to analyze theology as a science, Kuyper investigates the nature of science (Wissenschaft) and how theology relates to the other sciences. He was a prolific figure. Having left the pastorate to enter public life, Kuyper helped establish a political party, became a minister of parliament and later prime minister, founded a new denomination, helped found the Free University of Amsterdam, and wrote and published extensively. He lived through the Darwin era and late in life attended to evolution, drawing an important distinction between evolution as a worldview and evolution as a scientific theory. Not surprisingly, he is most well known as a public theologian.1 For Kuyper’s extensive writings, see Tjitze Kuipers, Abraham Kuyper: An Annotated Bibliography, 1857–2010, Brill’s Series in Church History 55 (Leiden: Brill, 2011).

1

36

Craig Bartholomew

The Enlightenment in Europe: Kuyper’s Context Kuyper had a front-row seat as the post-Enlightenment tradition took hold in biblical studies and theology in the Netherlands, and before his conversion, he appropriated much of it. However, once converted, and as his roles and responsibilities expanded, Kuyper came increasingly to see the antithesis between the Enlightenment worldview and the Calvinistic one.2 He saw clearly just how comprehensively the Enlightenment tradition had penetrated into all disciplines and all areas of society, not just theology. Kuyper recognized that a piecemeal, apologetic response to the comprehensive challenge of the Enlightenment would not work; the Enlightenment vision was comprehensive and required an equally comprehensive response from Christians. Later in life, when he gave his Stone Lectures on Calvinism at Princeton (1898), Kuyper described this comprehensive approach as a “worldview” (German: Weltanschauung).3 It is important to note that Kuyper never doubted the Enlightenment brought crucial insights and many gifts. Indeed, he is remarkably generous about the advances brought by it. In a discussion of common grace—a major theme in Kuyper’s thought4— and sin, for example, he asserts, Kuyper and the Scottish theologian James Orr (1844–1913) more or less simultaneously reached for the word worldview to give expression to the comprehensive nature of the gospel. Orr more helpfully uses the adjective Christian, but Kuyper was more significant in his outworking of the idea. See Craig G. Bartholomew, Contours of the Kuyperian Tradition: A Systematic Introduction (Downers Grove, IL: IVP Academic, 2017), chap. 4. 3 While Kuyper was capable of rigorous scholarship, his range in topics meant that he often painted with a broad brush, somewhat akin to Francis Schaeffer. In such cases his terms such as worldview and sphere sovereignty are always evocative but not always clearly defined. This can be frustrating to the scholar, but it positively calls for fresh engagement with Kuyper and a critical deepening of his work. It is especially among evangelicals that major work has been done on worldview in recent decades. See, for example, David Naugle, Worldview: The History of a Concept (Grand Rapids, MI: Eerdmans, 2002); and Michael W. Goheen and Craig G. Bartholomew, Living at the Crossroads: An Introduction to Christian Worldview (Grand Rapids, MI: Baker Academic, 2008). 4 Kuyper’s three volumes on common grace are now available in English: Abraham Kuyper, Common Grace: God’s Gifts for a Fallen World, vols. 1–3 (Bellingham, WA: Lexham Press, 2015–2020). The introduction for volume 2 is by Craig Bartholomew. There has been extensive discussion of Kuyper’s doctrine of common grace. For its background in Calvin and for the view that Herman Bavinck’s articulation avoids some of the problems in Kuyper’s, see Andrew McGowan, “Providence and Common Grace,” in The Providence of God: Deus Habet Consilium, ed. Francesca A. Murphy and Philip G. Ziegler (London: T&T Clark, 2009), 109-28. An equally important comparison to pursue is that between Kuyper and Herman Bavinck on the natural sciences and evolution. See Herman Bavinck, Essays on Religion, Science, and Society, ed. John Bolt, trans. Harry Boonstra and Gerrit Sheeres (Grand Rapids, MI: Baker Academic, 2008), chaps. 5–6. 2

Abraham Kuyper (1837–1920)

37

We very definitely may and must speak in this regard of God’s activity is immediately evident from the undeniable fact that in people like Plato and Aristotle, Kant and Darwin, stars of the first order have shined, geniuses of the highest caliber, people who expressed very profound ideas, even though they were not professing Christians. They did not have this genius from themselves, but received their talent from God who created them and equipped them for their intellectual labor.5

Another example is Kuyper’s approach to politics. His predecessor and mentor, Guillaume Groen van Prinsterer, wrote an important book, Lectures on Unbelief and Revolution, which, in line with Edmund Burke (1729–1797), was deeply critical of the French Revolution.6 It would be tempting to set a Christian approach in total opposition to that of the Revolution, but when Kuyper and his colleagues formed a political party they called it the Anti-Revolutionary Party. This expressed, of course, their opposition to the spirit of the French Revolution, but it also opposed the scorched-earth policy of revolutionaries, calling for reformation rather than revolution, thus recognizing the insights even of those they opposed. However, as the Enlightenment sought to reconstruct European culture from the ground up, Kuyper saw the challenge that it presented to the Christian faith at the most foundational level, and here he sought to challenge it strongly through developing a comprehensive Christian vision—or, as he would come to call it, a “worldview”—in a whole variety of ways in relation to the European and global challenges of the day. Understanding this context, as well as Kuyper’s contextual judgment, is crucial for understanding his engagement with science. God, Science (Wetenschap), and the Christian Mind Kuyper was neither a scientist nor a philosopher, but he could not fail to see just how central natural science was to the Enlightenment. His life overlapped with Charles Darwin’s (1809–1882), and by the end of Abraham Kuyper, Wisdom and Wonder: Common Grace in Science and Art, trans. Nelson D. Kloosterman (Grand Rapids, MI: Christian’s Library Press, 2011), 53-54. 6 See Harry Van Dyke, Groen van Prinsterer’s Lectures on Unbelief and Revolution (Jordan Station, ON: Wedge, 1989), which includes a translation of this work. 5

38

Craig Bartholomew

the nineteenth century, evolutionary thought was spreading like wildfire across Europe. As we will see, Kuyper took this development with full seriousness. Kuyper addresses science in at least four major places: in his work on common grace, in his Encyclopedia of Sacred Theology, in his Lectures on Calvinism, and in his October 1899 rectorial address on evolution at the Free University of Amsterdam. In this section we will focus on Kuyper’s complex, dynamic account of science, which provides the background for what might reasonably be considered his lasting legacy: worldview analysis (a term he first used in his Stone Lectures). It is important to note that as with the German word Wissenschaft, by science (Dutch, wetenschap) Kuyper is referring not just to the natural sciences but to all the university disciplines, and in this chapter we will use science this way and specify natural sciences otherwise. The distinction is more than semantic. Kuyper is calling our attention to the relationship between all the disciplines and what constitutes them as such. Although he was not a philosopher, when he came to write on theology, Kuyper needed to define science and why and how theology is one. Consequently, he ends up doing a considerable amount of foundational philosophical analysis of science in his Encyclopedia.7 We will begin with his comments on science and then explore his work on theology as a science, with special attention to how theology relates to the other sciences. Creation, “science,” common grace, palingenesis. Kuyper and his colleagues retrieved the doctrine of creation and made it central theologically and philosophically.8 Foundational theological concepts in Kuyper’s thought and view of science are creation, redemption (understood as grace restoring nature),9 “palingenesis,” special revelation, general revelation (or, as Kuyper refers to it, “common grace”), the image of God, and sin. Kuyper often develops his ideas from Greek words, and palingenesis is one such example. Since it crops up repeatedly in his disAbraham Kuyper, Encyclopedia of Sacred Theology: Its Principles, trans. Hendrik de Vries (New York: Charles Scribner’s Sons, 1898), 59. 8 Cf. Bruce R. Ashford and Craig G. Bartholomew, The Doctrine of Creation: A Constructive Kuyperian Approach (Downers Grove, IL: IVP Academic, 2020). 9 This is also a major emphasis of Herman Bavinck. 7

Abraham Kuyper (1837–1920)

39

cussion of science, it is worth explaining. In Titus 3:5, palingenesis means rebirth or regeneration in reference to individual conversion. In Matthew 19:28, however, palingenesis refers to the final renewal of all things. The word thus allowed Kuyper to stand firm on the imperative for individual regeneration and conversion while connecting this integrally to God’s purpose to renew all things. Conversion thus affects or at least should affect all we do and think, including science. For Kuyper the distinctive shape of the creation—what we might call its realism—has its origin in God’s thinking. He attends to John’s Logos doctrine (John 1) and argues that “the divine thinking must be embedded in all created things. Thus there can be nothing that fails to express, to incarnate, the revelation of the thought of God. A thought of God constitutes the core of the essence of things.”10 God’s thinking was gloriously original, as evoked in that rich wisdom text, Proverbs 8:22-31. When it comes to science, the crucial question becomes, Can humans gain access to God’s thoughts as they are embedded in the creation? Kuyper’s answer is a firm yes because humans are created in the image of God. The image includes wisdom, the capacity to know and understand the creation. Wisdom provides humans with an important “capacity . . . enabling them to pry loose from its shell, as it were, the thought of God that lies embedded and embodied in the creation.”11 For Kuyper, science is part of the cultural mandate (Gen 1:28), and, intriguingly, when he asserts the independence of science he observes that “if our human life had developed in its paradise situation, apart from sin, then science would have existed there. . . . Without sin there would be no state, and apart from sin there would have been no Christian church, but there would have been science.”12 This implies a dynamic view of the creation, the image of God, and their relation to one another. The creation is historically wired, and we image God as we are at work in his creation, excavating its potentials, understanding and knowing it, and developing it to God’s glory. This, in brief, is what the cultural mandate is all about. Kuyper, Wisdom and Wonder, 39. Kuyper, Wisdom and Wonder, 41. 12 Kuyper here expresses the typical Reformed view that the state is a consequence of the fall. In my view the Thomist approach is correct; the state is a normative part of the development of creation. Kuyper, Wisdom and Wonder, 35. 10 11

40

Craig Bartholomew

As a result, Kuyper’s view of science contains an interesting view of its history, a topic that has received major attention in recent decades. Science develops slowly and over time. It is never the work of one individual but rather a task that belongs to the whole human race, just as the whole human race is called corporately to image God. Because of the realism of creation—it has a discernible shape, and the God-given order of creation holds for everyone—science appears to arise by itself: “Rather, the entire temple is constructed without a human blueprint and without human agreement. It seems to arise by itself.”13 However, as one in the Reformed tradition, Kuyper insists that sin, and thus regeneration (palingenesis), affects not only the will but also the mind. He refers in this regard to the clear biblical distinction between true and false knowledge and argues that this applies as much to science as to any other human endeavor. “Sin is what lures and tempts people to place science outside of a relationship with God, thereby stealing science from God, and ultimately turning science against God.”14 The true ground of science is in the fear of the Lord, the beginning of wisdom (Prov 1:7). The difference between believing and unbelieving science, according to Kuyper, is not in the area being studied but “in the manner with which they [unbelievers] investigate, and in the principle from which people begin to investigate.”15 Thus, the mind or consciousness of the believing and unbelieving scientist differs. A Christian approach to science does not “keep itself busy with useless apologetics; it does not turn the great battle into a skirmish about one of the outworks, but immediately goes back to human consciousness, from which every man of science has to proceed as his consciousness. This consciousness, just on account of the abnormal character of things, is not the same in all.”16 Sin darkens the understanding or consciousness, and were it not for God’s common grace, science would have declined absolutely. However, this is not what we experience. The Greeks played a major role in the development of science, as did the Renaissance and the Enlightenment, Kuyper, Wisdom and Wonder, 45. Kuyper, Wisdom and Wonder, 51. 15 Kuyper, Wisdom and Wonder, 52. 16 Abraham Kuyper, Lectures on Calvinism (Grand Rapids, MI: Eerdmans, 1931), 136. Emphasis Kuyper’s, unless otherwise stated. 13 14

Abraham Kuyper (1837–1920)

41

often led by unbelievers. For Kuyper, this is because of God’s common grace. God sustains the creation in existence and holds back evil potentialities, thereby creating space for believers and unbelievers alike to develop creation. We will not understand science if we do not take both sin and common grace into account.17 Kuyper argues in his Lectures on Calvinism that science flows naturally from a Calvinistic perspective, and he makes much of Calvinism’s doctrine of common grace for unleashing this potential: A Calvinist who seeks God, does not for a moment think of limiting himself to theology and contemplation, leaving the other sciences, as of a lower character, in the hands of unbelievers; but on the contrary, looking upon it as his task to know God in all his works, he is conscious of having been called to fathom with all the energy of his intellect, things terrestrial as well as things celestial; to open to view both the order of creation, and the “common grace” of the God he adores, in nature and its wondrous character, in the production of human industry, in the life of mankind, in sociology and in the history of the human race. Thus you perceive how this dogma of “common grace” suddenly removed the interdict, under which secular life had laid bound, even at the peril of coming very near a reaction in favor of a one-sided love for these secular studies.18

This is not for a moment to suggest that Kuyper underplays the significance of a Christian perspective for science. Of palingenesis, for example, he notes that it is “a universal conception which dominates your whole person, and all of life about you; moreover, palingenesis exerts an influence not merely in your religious life, but equally in your ethical, aesthetical, and intellectual life.”19 The key question is how this works in practice. Kuyper has several answers to this question. First, believing and unbelieving scientists have different starting points or worldviews, and this influences one’s scientific work. Thus, second, “sin’s darkening lies in this, that we lost the gift of grasping the An important work on common grace and knowledge is J. J. Lee, “The Problem of Common Ground in Christian Apologetics: Towards an Integral Approach,” (PhD diss., North-West University, South Africa, 2014). Lee brings the Dutch tradition into dialogue with contemporary hermeneutics. 18 Kuyper, Lectures on Calvinism, 125. 19 Kuyper, Encyclopedia, 225. 17

42

Craig Bartholomew

true context, the proper coherence, the systematic integration of all things.”20 Third, Kuyper distinguishes between normalists and abnormalists when it comes to science. Unbelieving scientists approach the cosmos as we find it today as normal; they are normalists. Believing scientists, by contrast, approach the world today as fallen from its original goodness; they are abnormalists. For Kuyper this is where a major conflict lies: If it is normal, then it [the world] moves by means of an eternal evolution from its potencies to its ideal. But if the cosmos in its present condition is abnormal, then a disturbance has taken place in the past, and only a regenerating power can warrant it the final attainment of its goal. This, and no other is the principal antithesis, which separates the thinking minds in the domain of Science into two opposite battle-arrays.21

Abnormalists, for Kuyper, “do justice to relative evolution”22 but set original creation against naturalistic evolution, insist on humankind as a separate species made in the image of God, take sin seriously as deeply damaging to our original nature, and see the miraculous as the only way in which the abnormal can be restored. They find the norm for science not in the natural world but in the triune God. The conflict is not between faith and science, because all science proceeds on the basis of some faith: Notice that I do not speak of a conflict between faith and science. Such a conflict does not exist. Every science in a certain degree starts from faith, and, on the contrary, faith, which does not lead to science, is mistaken faith or superstition, but real, genuine faith it is not. Every science presupposes faith in self, in our self-consciousness; presupposes faith in the accurate working of our senses; presupposes faith in the correctness of the laws of thought; presupposes faith in something universal hidden behind the special phenomena; presupposes faith in life; and especially presupposes faith in the principles, from which we proceed; which signifies that all these indispensable axioms, needed in a productive scientific investigation, do not come to us by proof, but are established in our judgment by our inner conception and given with our self-consciousness.23 Kuyper, Wisdom and Wonder, 55. Kuyper, Lectures on Calvinism, 132. 22 Kuyper, Lectures on Calvinism, 132. 23 Kuyper, Lectures on Calvinism, 131. 20 21

Abraham Kuyper (1837–1920)

43

Thus, the conflict is between “two scientific systems, or, if you choose, two scientific elaborations . . . opposed to each other, each having its own faith.”24 All believing scientists must reckon with personal regeneration, its corresponding inspiration, the ultimate restoration of all things, and the manifestation of God’s power in and through miracles.25 For Kuyper science can only be truly one. Unbelieving science regards itself as the real thing, whereas believing science does likewise, so that there is an antithesis between the two.26 The two types of science are the same formally, but they have radically different starting points. “These two streams of science, therefore, which run in separate river-beds, do not in the least destroy the principle of the unity of science.”27 Although the starting points differ, “there is a very broad realm of investigation in which the differences exert no influence.”28 Kuyper identifies the following areas as fitting into this broad realm: (1) primary observation that restricts itself to numbers, weights and measures, (2) sense perception, “Any one who in the realm of visible things has observed and formulated something with entire accuracy, whatever it be, has rendered service to both groups,”29 and (3) logic, the formal process of thinking that is unaffected by sin. The natural sciences have the biggest substrate of these elements so that “it should be gratefully acknowledged that in the elementary parts of these studies there is a common realm, in which the difference between view- and starting-point does not enforce itself.”30 Kuyper distinguishes between the natural and spiritual sciences, and even in the latter he discerns a common realm in “outwardly observable facts,”31 referring to history and language. The higher the object studied, the more important palingenesis becomes. “With the faculty of Natural Philosophy, therefore, this antithesis makes itself least felt; a little more with the Medical; more strongly with the Philological; almost overwhelmingly with the Juridical; but most strongly of all with the Kuyper, Lectures on Calvinism, 133. Kuyper, Encyclopedia, 225. 26 Kuyper, Encyclopedia, 219-20. 27 Kuyper, Encyclopedia, 156. 28 Kuyper, Encyclopedia, 157. 29 Kuyper, Encyclopedia, 157. 30 Kuyper, Encyclopedia, 158. 31 Kuyper, Encyclopedia, 158. 24 25

44

Craig Bartholomew

Theological faculty.”32 Nevertheless, the common ground means that in all sciences the point of divergence between the two types of science must be closely attended to. Despite the common ground, for Kuyper, once a scientist starts making connections and moves to articulate general laws, differences emerge: Without a deep conviction of this unity, this stability and this order, science is unable to go beyond mere conjectures, and only when there is faith in the organic interconnection of the Universe, will there be also a possibility for science to ascend from the empirical investigation of the special phenomena to the general, and from the general to the law which rules over it, and from that law to the principle, which is dominant over all. The data, which are absolutely indispensable for all higher science, are at hand only under this supposition.33

Refreshingly, Kuyper argues that just as unbelieving science is not uniform, neither is palingenetic. The subjective element in scientists is unavoidable, and this creates divergence. Furthermore, it takes time for a Christian approach to science to develop, so differences are inevitable. “Friction, fermentation and conflict are the hall-mark of every experience of life on higher ground in this present dispensation, and from this the science of the palingenesis also effects no escape.”34 With conversion (palingenesis) comes a commitment to Scripture as God’s special revelation. How does this shape the scientist? The Revelation offered us in the Word of God gives us gold in the mine, and imposes upon us the obligation of mining it; and what is mined is of such a nature, that the subject as soon as he has been changed by palingenesis, assimilates it in his own way, and brings it in relation to the deepest impulse and entire inner disposition of his being. That this assimilation does not take place by means of the understanding only, can raise no objection, since it has been shown that naturalistic science also can make no advance without faith.35

Kuyper, Encyclopedia, 220. Kuyper, Lectures on Calvinism, 115-16. 34 Kuyper, Encyclopedia, 171. 35 Kuyper, Encyclopedia, 171-72. 32 33

Abraham Kuyper (1837–1920)

45

Since theology, according to Kuyper, “specializes” in palingenesis, our examination of Kuyper’s view of science would be incomplete without examining his view of theology. We now turn to this although in less detail.36 The science of theology and the theology of science. Kuyper defends the independence of science, and thus it behooves us to examine his view of theology as a science and how it relates to the other sciences. Kuyper attends closely to the difference between theology and other sciences. In the other sciences, the scientist stands above and over the object being investigated. However, when it comes to God, the “object” of the study of theology, this can never be the case: “Thinking man, taken as subject over against God as object, is a logical contradiction in terms.”37 For Kuyper, natural theology is of little consequence and help. He observes that palingenesis “applies to all faculties, but becomes more important in proportion as the part of the object which a given faculty is to investigate stands higher.”38 God stands highest by far, and this rules out natural theology: This whole matter assumes an entirely different phase, however, when palingenesis is taken as the starting-point. For then it ceases to be a problem whether there is a God; that the knowledge of God can be obtained is certain; and in the revelation which corresponds to this palingenesis there is presented of itself an objectum sui generis, which cannot be subserved under any of the other faculties; this impels the human mind to a very serious scientific investigation, which is of the utmost importance to practical life.39

Kuyper makes a distinction between God’s knowledge of himself (archetypal knowledge) and the knowledge of God revealed to humankind (ectypal knowledge). Theology attends to the ectypal knowledge of God since we can know God only if he reveals himself to us and in a form we can appropriate. “Only when that wondrous God will speak, can he listen. And thus the Theologian is absolutely dependent upon the pleasure of God, either to impart or not to impart knowledge of Himself.”40 For a fuller treatment, see Bartholomew, Contours of the Kuyperian Tradition, chap. 10. Kuyper, Encyclopedia, 214. 38 Kuyper, Encyclopedia, 220. 39 Kuyper, Encyclopedia, 223. 40 Kuyper, Encyclopedia, 251. 36 37

46

Craig Bartholomew

The imago Dei is the appropriate receptacle for God’s revelation. Kuyper develops this line of thought around his doctrine of the logos. There is the logos in the human person, and the Logos is the one through whom God made the world. We appropriate revelation through our logos. This is not to reduce theology to reason; it emerges from a living relationship with God. Indeed, Kuyper never equates Scripture with proposition. Today we would say that he is keenly alert to the great variety of speech acts in Scripture and in theology: “Christ does not argue, he declares; he does not demonstrate, he shows and illustrates; he does not analyze, but with enrapturing symbolism unveils the truth.”41 For Kuyper, the authoritative and determinative principle for theology is the Bible: “The material principium is the self-revelation of God to the sinner. From which principium the data have come forth in the Holy Scriptures, from which theology must be built up.”42 However, as he perceptively observes, while the “data” of theology is contained within the Bible, the Bible does not have the character of a theology textbook from which facts can easily be extracted and organized systematically. “The Holy Bible is, therefore, neither a law-book nor a catechism, but the documentation of a part of human life, and in that human life of a divine process.”43 This form of God’s revelation is relevant for the church of all times, that one holy, catholic and apostolic church: “God the Lord has spread one table for His entire Church, has given one organically connected revelation for all, . . . this one revelation . . . neither repeats nor continues itself . . . the witness of this one central revelation . . . lies for us in the Holy Scripture.”44 Kuyper is insightful in identifying different levels that set out the content of Scripture, from short confessional statements in Scripture, to creeds, to confessions, to scientific theology.45 “Wisdom” is given in Christ to all believers, but this is not the same as the “understanding” sought by theology, which “is that science which has the revealed knowledge of God

Kuyper, Encyclopedia, 287. Kuyper, Encyclopedia, 289. 43 Kuyper, Encyclopedia, 377. 44 Kuyper, Encyclopedia, 360. 45 Kuyper, Encyclopedia, 295-96. 41 42

Abraham Kuyper (1837–1920)

47

as the object of its investigation, and raises it to ‘understanding.’”46 Theology hereby yields scientific insight into the ectypal knowledge of God. “Science is called in, to introduce this knowledge of God, thus revealed, into our human thought”47 to the benefit of the church. So far so good. However, a crucial question arises: If science is independent but theology specializes in palingenesis,48 which relates to all of life, then what precisely is the relationship between theology and science? Kuyper deals with this in his Encyclopedia in a section titled “The Boundary of Theology in the Organism of Science.”49 He is clear that theology as a science communicates with all other disciplines. However, when its contributions are embraced by other disciplines, theology ceases to operate as itself: “As the theologian applies results furnished by logic, but is thereby no creator of logic himself, so the jurist, philologist, medicus and naturalist must deal with the results of theology, without themselves being thereby theologians.”50 It is the individual scientist and not the theologian who must work out how the data of God’s revelation and theology relate to his discipline. Kuyper is here speaking of the regenerate scientist but is also adamant in his defense of academic freedom: “He who investigates may render no obedience to any but the irresistible impulse of his own conviction.”51 Prior to the renewal of all things it is inevitable that conflicts will arise in and across all disciplines, including theology, and when it comes to the integration of data from nature and revelation, “in no particular should the naturalist . . . be impeded.”52 For Kuyper, “real advances will be made only when men who are themselves heart and soul alive to the efficacy of regeneration, at the same time devote all their powers of thought to these natural and historical studies, and so face these very conflicts.”53 By attending closely to the structure of science, Kuyper, in my view, provides an important example of retaining the normative authority of Kuyper, Encyclopedia, 299. Kuyper, Encyclopedia, 327. 48 See Kuyper, Wisdom and Wonder, 33-45. 49 Kuyper, Encyclopedia, 605-15. 50 Kuyper, Encyclopedia, 606. 51 Kuyper, Encyclopedia, 606. 52 Kuyper, Encyclopedia, 607. 53 Kuyper, Encyclopedia, 608. 46 47

48

Craig Bartholomew

Scripture for all of life while avoiding the dangers of both biblicism, in which, for example, one reads theories of natural science straight off Genesis 1:1–2:4, and dualism, in which science and faith operate separately and autonomously. Kuyper articulates a humble role for theology in relation to the other disciplines and rejects the view of theology as queen of the sciences. He also argues that we need to clearly distinguish Christian theology from Christian philosophy. We need the latter, and for Kuyper, its role is to organize into a unity the knowledge gleaned from the different sciences. Theology and the Natural Sciences: Evolution Kuyper’s discussion of evolution is surprisingly detailed. I will focus on his rectorial address, given at the Free University of Amsterdam on ­October 20, 1899, two months after his beloved wife had died.54 Nowadays it is astonishing to think of this as a presidential address at a university. The English translation of the full text comes in at just under eighteen thousand words! What is equally astonishing is the breadth and depth of Kuyper’s reading. Bratt notes that the published version includes some forty footnotes, but Kuyper’s reading manuscript indicates he had planned some seventy more.55 Because of his wife’s death, he was unable to include these. In his footnotes Kuyper references some twenty-seven authors, often with multiple works. In true European fashion, Kuyper reads across German (in particular), French, English (Darwin), and Dutch sources. The historical and ongoing significance of many of these figures remains. Most books he references were published around the time he wrote, and authors include such luminaries as George John Romanes (1848–1894), an evolutionary biologist and physiologist who was a close friend of Charles Darwin; Friedrich Weismann (1834–1914), described as the most well-known evolutionist after Darwin in the nineteenth century; and Herbert Spencer The translation of the full text is available online at https://www.asa3.org/ASA/resources/ Kuyper.html. See also James D. Bratt, ed., Abraham Kuyper: A Centennial Reader (Grand Rapids, MI: Eerdmans, 1998). Bratt’s version represents 90 percent of the original text (pp. 403-40). References that follow are from Bratt’s edition. 55 Bratt, Abraham Kuyper, 404. 54

Abraham Kuyper (1837–1920)

49

(1820–1903),56 who originated the term “survival of the fittest,”57 which was adopted by Darwin, and developed a comprehensive theory of evolution embracing all aspects of life. Toward the close of the nineteenth century, Spencer was one of the most famous European intellectuals. He applied evolution to religion and ethics,58 arguing for a utilitarian view of the latter. He is well known for his disturbing social Darwinism, which applied the law of the survival of the fittest to societal development, affirming racist views in the process. In his Social Statistics, for example, he writes, The forces which are working out the great scheme of perfect happiness, taking no account of incidental suffering, exterminate such sections of mankind as stand in their way, with the same sternness that they exterminate beasts of prey and herds of useless ruminants. Be he human being, or be he brute, the hindrance must be got rid of. Just as the savage has taken the place of lower creatures, so must he, if he have remained too long a savage, give place to his superior.59

Ernst Haeckel (1834–1919) was a German zoologist and biologist who discovered and named thousands of species. His exquisite drawings that enhanced his work remain available and popular today. Haeckel was deeply influenced by Darwin’s On the Origin of Species when he read it in 1864, and, pushing aside Darwin’s cautions, he developed a wideranging and comprehensive evolutionary philosophy and history of life.60 See Mark Francis, Herbert Spencer and the Invention of Modern Life (London: Routledge, 2014). Francis says, “I claim that Spencer was not Darwinian, either in his biological writing or in his account of human evolution. My reasons for distinguishing between Spencer’s theories and Darwinism are: (i) Spencer’s evolutionary theory did not focus on species change; (ii) Spencer’s faith in progressive evolution did not draw on natural selection or competition; and (iii) Spencer did not accept that modern individuals and societies would continue to make progress through struggle for survival. My insistence on the non-Darwinian quality of Spencer’s ideas does not refer to anything intrinsically interesting about Spencer’s evolutionary stance” (2-3). 57 Francis, Herbert Spencer, 3, explains that in a conversation before Darwin published his Origin, “After Darwin had explained his theory of natural selection, Spencer quipped that it might as well be called ‘survival of the fittest.’ Subsequently, Darwin adopted this phrase as describing evolutionary theory while its originator did not.” 58 See Herbert Spencer, The Data of Ethics (London: Routledge, 2011). 59 Herbert Spencer, Social Statistics: Or the Conditions Essential to Human Happiness Specified, and the First of Them Developed (London: John Chapman, 1851), 416. 60 Many of Haeckel’s works are readily available in English. Secondary sources I have consulted include Robert J. Richards, The Tragic Sense of Life: Ernst Haeckel and the Struggle over Evolutionary Thought (Chicago: University of Chicago Press, 2008); Sander Gliboff, H. G. Bronn, Ernst 56

50

Craig Bartholomew

Haeckel’s name occurs forty-nine times in Kuyper’s “Evolution,” indicating his role as Kuyper’s major dialogue partner. For Kuyper, evolutionists have been brilliant in uncovering new facts for which we must be grateful. Evolutionary theory is to be welcomed as progress in relation to the detailed empiricism and shroud of the unknowable that had hung over natural philosophy. Evolution has been rigorous in its detailed observation of nature. Previously it was as though we had only been looking at the face of a clock; evolution has opened up the underlying mechanism to enable us to take a very close look. It has provided an impetus to ontogenetic and morphological studies; it has uncovered a unity of design in organic life. Kuyper is quite open to God having used evolution as his means of creating. He argues, Of course, it is an entirely different question . . . whether religion as such permits a spontaneous unfolding of the species in organic life from the cytode or nuclear cell. This question must be answered affirmatively, without reservation. We will not force our style upon the Chief Architect of the universe. If he is to be the Architect not in name only but in reality, He will also be supreme in the choice of style. Had it thus pleased God not to create the species, but to have one species emerge from another by enabling a preceding species to produce a higher following species, Creation would still be no less miraculous.61

Thus, Kuyper is clear that we must not deny but rather welcome the facts that evolution has uncovered. However, facts always require explanation— as modern philosophy of science has noted, theory is underdetermined by facts—and Kuyper vehemently opposes the extrapolation from the facts of evolution to a grand theory of Evolution (capital E) that explains everything. The Enlightenment tradition had now found a unified principle in evolution, which it thinks can explain everything. For Kuyper this is disastrous and dangerous, and he rejects the view of mechanistic action as driving evolution exempt of any spirit that forms and rules it. He finds Haeckel, and the Origins of German Darwinism: A Study in Translation and Transformation (Cambridge, MA: MIT Press, 2008); and Andrea Wulf, The Invention of Nature: The Adventures of Alexander Von Humboldt, The Lost Hero of Science (London: John Murray, 2015). 61 Abraham Kuyper, “Evolution,” in Bratt, Abraham Kuyper, 436-37.

Abraham Kuyper (1837–1920)

51

Evolution reductionistic, as though the history of the world can be reduced to a physical-chemical process. That scientists were extrapolating evolution into Evolution as a grand theory is clear from Haeckel’s work, for example. Kuyper does not explore Haeckel’s life story, but it is fascinating. Haeckel had a conventional upbringing in the German Evangelical Church and wrestled with how to relate his faith to science. In November 1861 he immersed himself in Darwin’s Origin, and “from that fertile womb he emerged newly born for Darwin’s theory, and the zeal of his conviction never cooled through the later days.”62 However, a profoundly formative experience for Haeckel was the death of his wife, Anna, in 1864, on the day he turned thirty. As Robert Richards notes, “The death of his first wife severed the loose threads still holding him to formal observance. The power of that death, his obsession with a life that might have been, and the dark feeling of love forever lost—these haunting remains drove him to find a more enduring and rational substitute for orthodox religion in Goethean nature and Darwinian evolution.”63 In a note he wrote about ten years after Anna’s death, Haeckel acknowledged that her death “destroyed with one blow all the remains of my earlier dualistic worldview.”64 In his espousal of evolution and in his development of it into a comprehensive account of all of life, Haeckel intentionally targeted orthodox Christianity: The antagonism between conservative religion and evolutionary theory, brought to incandescence at the turn of the century and burning still brightly in our own time, can be attributed, in large part, to Haeckel’s fierce broadsides launched against orthodoxy in his popular books and lectures. These attacks and reactions to them escalated to a new level during the period from 1880 to his death in 1919.65

It is precisely the sort of advocacy of Evolution that Kuyper opposes. He observes a hierarchy in the universe, which leads him to question, “Must all higher organized life be pulled down to the spheres of lower Richards, Tragic Sense of Life, 68. Richards, Tragic Sense of Life, 343. 64 Richards, Tragic Sense of Life, 107. In his struggles over science and religion, Haeckel had found comfort in Schleiermacher’s separation of the two spheres of knowledge and religion. On Schleiermacher’s two sphere’s approach, see Sonderegger, chapter five of this volume. 65 Richards, Tragic Sense of Life, 344. 62 63

52

Craig Bartholomew

organic life, or must the lower be subsumed under the higher?”66 Kuyper is adamant that absolutely nothing in nature is alien to us; however, the spiritual sphere has an autonomous character. He finds Evolution’s attempt to account for aesthetics, ethics, and religion sorely wanting: The moral ideal, the moral world order, the moral law that governs us, the sense of duty that binds us to that law, and the Holy One who gives us the law all fall away, and with these basic ideas we lose the correlate ideas of sin, guilt and repentance, redemption and atonement. Thus Evolution robs Ethics of its entire subject [matter].67

Kuyper also associates Nietzsche with Evolution. He asserts that in opposition to the view that the strong must tread down the weak, we set the merciful Christ. In opposition to an undirected mechanistic evolution, we set God who works all things according to the counsel of his will (Eph 1:11). In opposition to death’s annihilation, we set a coming judgment and eternal glory. When Kuyper gave his address on October 20, 1889, his wife had died two months earlier, and doubtless there was much in his personal life to make him melancholic. His sense of the nineteenth century as ending on a bad note is also intriguing, considering what lay in wait for Europe. “The call for bread and circuses, the dissolution of the marriage bond, and so much more create the perception that the decline and fall of the Roman empire is being repeated on an even more terrible scale in our touted age.”68 For Kuyper, the major challenge of Evolution is not only that the chimpanzee is touted as our ancestor but that mercy as a primary virtue is lost. In his view, Evolution as a metanarrative ultimately leads to nihilism. Although the theory of evolution has developed in major ways since Kuyper, much of what he says remains relevant. Charlesworth, for example, notes that “although there are still many unsolved problems, biologists are convinced that even the most complicated features of living creatures, such as human consciousness, reflect the operation of chemical Kuyper, “Evolution,” 431. Kuyper, “Evolution,” 434. 68 Kuyper, “Evolution,” 410. 66 67

Abraham Kuyper (1837–1920)

53

and physical processes that are accessible to scientific analysis.”69 Here we find the same reductionism of everything to physical-chemical processes that Kuyper opposed. This statement also reveals the tendency, so criticized by Kuyper, to extend evolution to Evolution. Again, “Evolutionary ideas provide a set of natural processes that can explain the vast diversity of living species, and the characteristics that make them so well adapted to their environment, without any appeal to a mind that designed them.”70 Kuyper rightly stresses the difference between natural selection and Christ’s emphasis on mercy. Charlesworth acknowledges that “evolution is also pitiless. Selection acts to hone the hunting skills and weapons of predators, without regard to the feelings of their prey.”71 One cannot help but discern a faith in evolution when Charlesworth concludes that “an understanding of evolution can teach us our true place in nature, as part of the immense array of living forms which the impersonal forces of evolution have produced.”72 There are certainly areas where Kuyper should not be followed today. I find, for example, his Logos doctrine and his view that we can find God’s thoughts in the creation overly intellectual. Although he was not a natural scientist, Kuyper saw the challenges of the sciences—and not least, the natural sciences—as European culture was being reconstructed, and he recognized the need to respond. We are in a similar situation, in which the natural sciences are too often regarded as the final word, including on topics way beyond their remit.73 Kuyper, I argue, asks the right questions: What is science? How does our worldview impact science? How do sin and redemption affect science? We need to attend afresh to such questions today. In Conversation with Abraham Kuyper It should be clear that Kuyper was a fascinating and insightful interlocuter. Of course, science and theology have developed in all sorts of Brian Charlesworth, Evolution: A Very Short Introduction, 2nd ed., Very Short Introductions (Oxford: Oxford University Press, 2017), 4-5, emphasis added. 70 Kuyper, “Evolution,” 5. 71 Kuyper, “Evolution,” 132. 72 Kuyper, “Evolution,” 133. 73 Cf. James Davison Hunter and Paul Nedelisky’s discussion of the natural sciences’ proclivity to overreach in Science and the Common Good: The Tragic Quest for the Foundations of Morality (New Haven, CT: Yale University Press, 2018), 119-37. 69

54

Craig Bartholomew

ways since Kuyper, but there remains a great deal we can learn from him. Where I find Kuyper always helpful is his asking the right questions and his rigor in pursuing them. The natural sciences are performed by humans in the world. Thus, what we understand by the natural sciences will be shaped by our worldview and our view of what it means to be human. Always, therefore, the natural sciences will work with some ontology and some anthropology. Of course, they will also work with some sense of how to know the world truly, and thus with some epistemology. None of these building blocks are neutral, and scholars differ profoundly on these issues even when, or especially when, they are unconscious of them. Quite naturally, scholars in all disciplines, including the natural sciences, generally get on with work in their chosen specialty. However, as Charles Taylor notes, within any discipline, including the natural sciences, you can ask deeper and deeper questions until you reach the most foundational ones, and that is philosophy. With Kuyper I would argue that at an even deeper level we find worldview. Kuyper’s genius is to attend closely to the foundations of science. He rightly notes that how we understand the shape of reality, epistemology, and our view of the human person will deeply shape our view of science. As a Christian, Kuyper approaches the world as God’s good but fallen creation in the process of being redeemed. For Kuyper, science is possible because of the way God has ordered the world and because humans are made in his image. This remains a powerful argument. Alvin Plantinga has, for example, comparably argued that there is nothing in naturalistic evolution which would lead us to be sure that human cognitive abilities have evolved in such a way as to know the world truly. They would have evolved to enhance survival, but this need not include a capacity to know the world correctly. Plantinga argues that Darwin himself had such doubts.74 Kuyper explores vital questions such as the difference it makes to science when the world is understood as God’s creation, and the effect of sin and redemption on science. This remains a crucial issue requiring rigorous thought. Like most American evangelicals of his time, Kuyper thought of the natural sciences as a common human endeavor in which Christian and non-Christian could work on the same epistemic foundations. Alvin Plantinga, “Evolution vs. Naturalism,” Books and Culture, July–August 2008.

74

Abraham Kuyper (1837–1920)

55

However, as Michael Buckley points out in his seminal At the Origins of Modern Atheism,75 the major mistake Christians made in relation to the Enlightenment was to concede epistemic foundations in the hope that they could work on them toward truth that would agree with Christian truth. George Marsden helps us to understand why American evangelicals took the view they did, but, as he notes, in the process they inadvertently kept the door open to the radical Enlightenment until it was too late.76 Writing in 1991, Marsden concluded that, At the moment the issues in the intraevangelical debate are far from resolved. The Warfieldians still look back, in effect, to the days of evangelical hegemony in the emerging scientific culture and still look forward to a time when true science and true Christianity will be compellingly synthesized. The Kuyperians, on the other hand, are frankly more pluralistic in their view of the human scientific community. . . . Much of the confusion . . . stems from the lack of resolution . . . of these intellectual issues.77

This assessment seems to me as true now as it was in 1991. However, life does not stand still, and in my view, the Princetonian approach, which remains alive and well among evangelicals, easily polarizes evangelical opinion between an uncritical embrace of evolution—science is science, after all!—or an equally uncritical opposition to it in the name of creationism. What is too often lacking or not even understood is that all natural science operates out of a paradigm or a worldview, and extremely hard work has to be done on the ways in which different perspectives or worldviews and philosophies shape the theories that emerge so that Christians can adopt the genuine insights without taking on the ideological baggage. Kuyper is more helpful than Warfield in pointing us in this direction. However, even Kuyper at points seems in danger of conceding too much common ground.78 How so? Common grace is a major tool in Michael J. Buckley, At the Origins of Modern Atheism (New Haven, CT: Yale University Press, 1990). See Jonathan Israel’s Radical Enlightenment: Philosophy and the Making of Modernity, 1650–1750 (Oxford: Oxford University Press, 2001) for the case that there is a greater coherence to the Enlightenment than is often realized and that the dominant group is that of the radicals. 77 George M. Marsden, Understanding Fundamentalism and Evangelicalism (Grand Rapids, MI: Eerdmans, 1991), 151-52. 78 Herman Dooyeweerd’s (1894–1977) philosophical work is more penetrating in this respect. 75

76

56

Craig Bartholomew

Kuyper’s toolbox, and it enables us to see how unbelievers produce excellent work and profound insights about creation. As we have seen, Kuyper is adamant that wherever facts are identified, we must take them seriously. Certainly Kuyper is right that the Christian faith implies some form of philosophical realism—namely, that creation has a discernible shape that can be known. However, Kuyper’s repeated emphasis on facts and on perception being unaffected by sin bears correction by recent philosophy of science. A. F. Chalmers notes that “we have seen that facts adequate for science are by no means straightforwardly given but have to be practically constructed, are in some important senses dependent on the knowledge that they presuppose . . . and are subject to improvement and replacement.”79 Thus, the subjective factors in science go deeper than even Kuyper proposes. One way to express this is that God’s created order imposes itself on all of us and is common to all of us, but how we go about knowing it (epistemology) is shaped by our foundational presuppositions and worldview. Thus, Kuyper’s very useful distinction between Evolution and evolution may in the end be too neat and tidy. Natural scientific theory is underdetermined by the facts, and there is always a hermeneutic at work identifying and assembling the facts identified into a whole. This hermeneutic can never be taken for granted but needs to be interrogated just as much as any other aspect of natural scientific theory. If Kuyper were alive today, I would love to hear his views of the Intelligent Design movement. One thing I am sure of: he would read and reflect on all the literature and explore it critically, just as he did in the late-nineteenth century as evolution and Evolution swept across Europe. On the whole I find Kuyper’s approach to theology helpful. Indeed, I side with him and Herman Bavinck against Herman Dooyeweerd, who sees theology as only one of the special sciences, with philosophy as the science of the sciences. Theology reflects on God’s special revelation (namely, Scripture) and seeks to systematize its teaching for today. Scripture is authoritative for all of life, and thus theology must be taken into account by other disciplines. In my view both theology and A. F. Chalmers, What Is This Thing Called Science?, 3rd ed. (Indianapolis, IN: Hackett, 1999), 58.

79

Abraham Kuyper (1837–1920)

57

philosophy are foundational disciplines with major implications for every other discipline, an approach of the sort we find in Augustine. Rather than organizing the fruits of the sciences into a unity, in my view philosophy, like theology, has a far more foundational role and is already at work in any and every science. Thus, the dialogue between Christian philosophy, theology, and the natural sciences is essential. One takes comfort from Kuyper’s view that there will be a pluriformity among believing scientists, and that agreement will not be quickly achieved. What cannot be accepted is for natural scientists to regard their terrain as sealed off from the effects of palingenesis and thus Scripture and theology. Kuyper’s strong assertion of the independence of science is vulnerable to such a move, but this is, I think, to misunderstand it and should be resisted. I find Kuyper’s doctrine of sphere sovereignty very insightful,80 but it has to be balanced by his view of sphere universality and should not be used to justify an autonomous view of science. There should certainly be a healthy Christian academic freedom, but it is hard to see how it could be Christian if, for example, the catholic creeds are not taken as normative. There is thus room for a critical reexamination of Kuyper’s view of the independence of science. I noted above that Kuyper’s analysis of the structure of science is an example of retaining Scripture as normative for all of life, including the natural sciences, while avoiding biblicism and dualism. However, it is an example. The relationship between Scripture and the natural sciences continues to be a major point of conflict, and major work needs to be done in this area. With Kuyper we need to attend to the structure of theorizing in the natural sciences so that the whole picture comes into view. Where recent philosophy of language and philosophy of science have developed in fascinating ways is in the area of metaphor. Traditionally it was thought that scientific language should erase metaphor. Now we are coming to see that metaphors are the building blocks of theory.81 Scripture too is chock full of metaphors, and it is in the area of See Bartholomew, Contours of the Kuyperian Tradition, chap. 5. See, for example, Michael A. Arbib and Mary B. Hesse, The Construction of Reality (Cambridge: Cambridge University Press, 1986); and M. Elaine Botha, Metaphor and Its Moorings: Studies in the Grounding of Metaphorical Meaning (New York: Peter Lang, 2007).

80 81

58

Craig Bartholomew

networks of metaphors that hope may lie for a fresh examination of the relationship between Scripture, theology, and the natural sciences. This is a project I hope to pursue. Recommended Reading Primary Works Kuyper, Abraham. Encyclopedia of Sacred Theology: Its Principles. New York: Charles Scribner’s Sons, 1898. ———. “Evolution.” In Abraham Kuyper: A Centennial Reader, edited by James D. Bratt, 403-40. Grand Rapids, MI: Eerdmans, 1998. Note that this entire volume is a gold mine of Kuyper’s writings. ———. Lectures on Calvinism. Grand Rapids, MI: Eerdmans, 1931. ———. Wisdom and Wonder: Common Grace in Science and Art. Grand Rapids, MI: Christian’s Library Press, 2011. allofliferedeemed.co.uk/kuyper.htm contains a wealth of primary sources by Kuyper.

Secondary Works Bartholomew, Craig G. Contours of the Kuyperian Tradition: A Systematic Introduction. Downers Grove, IL: IVP Academic, 2017. Bratt, James D. Abraham Kuyper: Modern Calvinist, Christian Democrat. Grand Rapids, MI: Eerdmans, 2013. Heslam, Peter S. “Architects of Evangelical Intellectual Thought: Abraham Kuyper and Benjamin Warfield.” Themelios 24, no. 2 (1999): 3-20. ———. Creating a Christian Worldview: Abraham Kuyper’s Lectures on Calvinism. Grand Rapids, MI: Eerdmans, 1998. See chap. 7. Klapwijk, Jacob. “Kuyper on Science, Theology and University.” Philosophia Reformata 78, no. 1 (2013): 18-46. Marsden, George M. Understanding Fundamentalism and Evangelicalism. Grand Rapids, MI: Eerdmans, 1991. See chaps. 5 and 6.

3 B. B. WA R F I E L D ( 1 8 5 1 – 192 1 ) Evolution, Human Origins, and the Development of Theology Bradley J. Gu ndl ach Benjamin Breckinridge Warfield came from a Kentucky family distinguished in politics and religion. He was educated at Princeton College and Princeton Seminary, taught New Testament at Western Theological Seminary, and established himself as an expert on textual criticism. In 1887 he succeeded A. A. Hodge as the chair of theology at Princeton, a position he held until his death. After a difficult coeditorship of The Presbyterian Review with Charles Briggs, he founded and edited the highly respected Presbyterian and Reformed Review to champion confessional Calvinism and Christian supernaturalism through up-to-date scholarship. Next to Charles Hodge, Warfield was the most important exponent of the Princeton theology, which dominated American Presbyterianism in the nineteenth century and on which a wide spectrum of evangelicals and fundamentalists drew selectively in the twentieth century. His name, more than any other, is associated with the doctrine of the inerrancy of Scripture in the original autographs, a doctrine he refined and defended with massive erudition. He also addressed the questions of science, history, and Scripture that occupied so many in his generation, offering a view expressly opposed to that of Friedrich Schleiermacher’s limitation of scriptural authority to matters of faith and practice.

W

hat did Benjamin Breckinridge Warfield actually hold concerning evolution and human origins? Lately there has been disagreement on this question (or rather these questions). While Warfield has long been identified as a “theistic evolutionist,” offering sound precedent for evangelicals to take that position, Fred Zaspel has countered with the claim that Warfield never accepted evolution and in fact uttered many

60

Bradley J. Gundlach

damning criticisms of it.1 At the 2017 meeting of the Evangelical Theological Society, Zaspel’s work was featured in the rollout of a new multidisciplinary polemical volume against theistic evolution.2 The session attracted a huge audience. Interpretations of Warfield’s views on evolution and human origins fall into three categories, with some spectrum of opinion within. The Warfield-as-evolutionist camp includes J. I. Packer, Mark Noll and David Livingstone, and Peter Enns. “I recall that B. B. Warfield was a theistic evolutionist,” Packer wrote in 1978. “If on this count I am not an evangelical, then neither was he.”3 Noll and Livingstone, in their fine collection of Warfield’s writings on the subject, call it “one of the best-kept secrets in American intellectual history” that “B. B. Warfield, the ablest modern defender of the theologically conservative doctrine of the inerrancy of the Bible, was also an evolutionist.”4 This strong statement, admittedly designed to jolt discussion into more fruitful channels, is nuanced soon thereafter, when Noll and Livingstone describe Warfield as “a cautious, discriminating, but entirely candid proponent of the possibility that evolution might offer the best way to understand the natural history of the earth and of humankind.”5 Their inclusion of humankind here can be misleading, as I will argue below, but their emphasis on Warfield’s caution and discrimination is exactly right. Pete Enns provides us with a good example of a less careful appeal to Warfield, declaring on the BioLogos website that Warfield, “as is well known, accepted evolution as giving the proper scientific account of human origins.”6 See, for example, Fred G. Zaspel, “B. B. Warfield on Creation and Evolution,” Themelios 35, no. 2 (2010): 198-211, http://themelios.thegospelcoalition.org/article/b.-b.-warfield-on-creation -and-evolution. 2 While the book has thirty-one chapters, Zaspel’s was one of the several chosen for presentation at the conference. Fred Zaspel, “Additional Note: B. B. Warfield Did Not Endorse Theistic Evolution as It Is Understood Today,” in Theistic Evolution: A Scientific, Philosophical, and Theological Critique, ed. J. P. Moreland et al. (Wheaton, IL: Crossway, 2017), 953-72. 3 J. I. Packer, The Evangelical Anglican Identity Problem (Oxford: Latimer House, 1978), 5, quoted in Alister McGrath, The Foundations of Dialogue in Science and Religion (Malden, MA: Blackwell, 1998), 130. 4 Mark A. Noll and David N. Livingstone, eds., Evolution, Science, and Scripture: Selected Writings (Grand Rapids, MI: Baker Books, 2000), 14. 5 Warfield, Evolution, Science, and Scripture, 15. 6 Pete Enns, “Evolution and Our Theological Traditions: Calvinism” BioLogos blog, March 25, 2011, https://biologos.org/articles/evolution-and-our-theological-traditions-calvinism. 1

B. B. Warfield (1851–1921)

61

Agreeing with this identification of Warfield as an evolutionist is the late Henry Morris, leader of the modern creationist movement. For him, Warfield’s example serves evangelicals not as sound precedent but as a warning—Warfield was guilty of “pervasive theological apostasy.” He betrayed the doctrine of biblical authority by caving in to evolutionism, and anyone who followed him would be put on the slippery slope.7 Fundamentalist suspicion toward Warfield on evolution marks the second category of interpretations. Zaspel now opens a third view: he sees Warfield as an evangelical and inerrantist who indeed modeled openness to scientific theories of evolution but whose example in not accepting them we should follow. Zaspel portrays Warfield as moving away from his earlier allowances for evolution by the end of his career and as never having accepted it outright once he started thinking theologically about the issue. Zaspel builds his case on multiple citations from Warfield’s writings, especially on Warfield’s statement in 1916 that though he came to college as already “a Darwinian of the purest water,” he “fell way” from it around the age of thirty.8 We have, then, at least three different takes on Warfield and evolution (not counting my own)9: Warfield as (1) the positive example of a sound evangelical evolutionist; (2) the negative example of an erring evangelical evolutionist, sound on most doctrine but insidious in his acceptance of evolution; and (3) the positive example of a sound evangelical who considered evolutionism but ultimately rejected it. Warfield’s example matters so much because of his formative role in the modern evangelical formulation and defense of biblical inerrancy. What, then, did Warfield actually hold concerning the evolution of species, especially concerning the origin of humankind? I will front-load my conclusions here, then substantiate them by appealing to the most important of his publications on the subject as well as to other evidence Henry M. Morris, A History of Modern Creationism (San Diego, CA: Master Book Publishers, 1984), 39, quoted in McGrath, Foundations of Dialogue, 130. 8 Fred Zaspel, The Theology of B. B. Warfield: A Systematic Summary (Wheaton, IL: Crossway, 2010), 386. 9 Bradley J. Gundlach, Process and Providence: The Evolution Question at Princeton, 1845–1929 (Grand Rapids, MI: Eerdmans, 2013). The present essay builds on my earlier work, going somewhat beyond it. 7

62

Bradley J. Gundlach

not usually brought to bear on the question. Along the way I will turn our attention to ways in which developmental thinking pervaded Warfield’s theology outside the question of biological origins. Warfield clarified and developed his position on evolution throughout his career, but its overall shape is clear and consistent. His position includes the following theses: 1. The earth may well be very old—and the question even of the antiquity of man is of no real consequence to the theologian. 2. The genealogies in Genesis were not intended to yield a chronology. 3. Rather, their purpose is to impress us with “the greatness of those grand men of old, towering, as they did, in strength and endurance above all that the world has since seen.”10 Thus, Warfield took as fact the extreme longevity of the antediluvians. 4. Species may have come from other species by means of descent under the providential government of God. 5. But God created the original world-stuff ex nihilo and intruded supernatural power to create life and to make man—perhaps at many other steps too. 6. God’s providential superintendence of the world is not limited to the use of natural laws. 7. Theistic evolution is unbiblical and insufficient for the Christian. 8. Adam may have had a brute ancestry but was not himself the product of evolution. 9. The biblical doctrines of God, man, and sin contradict the theory of human evolution. 10. Creation and evolution are mutually exclusive concepts, for evolution is providence. The first thing to observe about this list is how it defies categorization. It may seem even to lack coherence. Warfield allows brute ancestry for B. B. Warfield, “The Manner and Time of Man’s Origin,” in Evolution, Science, and Scripture, 220-21. The article originally appeared in The Bible Student, n. s. 8, no. 5 (November 1903): 241-52.

10

B. B. Warfield (1851–1921)

63

humankind but denies human evolution. He takes the ages of the antediluvians literally but says they are not intended for constructing a chronology. He allows, even holds as likely, the transmutation of species yet says it is wrong to speak of God “creating by evolution.” No wonder there is disagreement as to Warfield’s embrace of evolutionism. A Brief Sketch of Warfield’s Historical Moment To understand Warfield’s views on evolution and theology, we must consider his situation in time and place. He came of age at just the time when Darwin’s work was gaining acceptance in scientific circles but encountering opposition in the church. He grew up in a family of Kentucky Presbyterians distinguished in religion and politics, the Breckinridges, on his mother’s side. His father and grandfather Warfield were pioneering breeders of shorthorn cattle, so young Ben Warfield came to the evolution question from actual experience of the effects of selection in breeding.11 As he later recalled, when he went to college at Princeton at the age of sixteen in 1868, “I was already a Darwinian of the purest water . . . and knew my Origin of Species and Animals and Plants Under Domestication almost from A to Izard.”12 That autumn James McCosh arrived as Princeton’s dynamic new president, the first American religious leader to come out publicly for the doctrine of evolution. McCosh’s Christianity and Positivism (1871) embraced the transmutation theory while at the same time forcefully countering the metaphysical and epistemological dangers that tended to accompany the theory. Warfield called McCosh “distinctly the most inspiring force” he encountered in his college days— not because McCosh convinced him of evolution but presumably because McCosh modeled a serious engagement of Christian conviction with natural science, confidently showing that one could follow modern learning and still believe the Bible’s this-worldly claims accessible to historical and scientific scrutiny, not just its spiritual teachings. In this way Warfield was launched on a trajectory fundamentally different from the David Livingstone first made this point. See David N. Livingstone and Mark A. Noll, “B. B. Warfield (1851–1921): A Biblical Inerrantist as Evolutionist,” Isis 91 (2000): 291-93. 12 B. B. Warfield, “Personal Recollections of Princeton Undergraduate Life: IV. The Coming of Dr. McCosh,” Princeton Alumni Weekly 16, no. 28 (April 19, 1916): 652. Notice that the latter title was published that very year. 11

64

Bradley J. Gundlach

followers of Schleiermacher.13 McCosh’s writings “averted a disastrous war between science and faith,” said George Macloskie, professor of natural history, “and in ‘his’ college, men have studied Biology without discarding their religion.”14 Warfield was a student at Princeton Seminary when Charles Hodge wrote What Is Darwinism? and concluded, in the most-quoted anti-Darwinian line ever, “It is atheism.”15 Warfield would follow Hodge in the view that Darwinism itself—evolution by natural selection with the express denial of divine design—was indeed atheistic, though other evolutionisms were not.16 By the time Warfield came back to Princeton Seminary as professor of systematic theology in 1887, evolutionism had taken the field in science and many religious leaders had joined McCosh in seeing it as “God’s method of creation.” But change over time by natural causes was becoming a fascination far beyond the life sciences, and the triumphs of “Darwinism” (as a catchall for any and all transmutation theories) now lent scientific prestige to older theories about the origins of Scripture and to newer theories about the “progress” of Christianity in the rejection or recasting of old orthodoxies. Eventually the controversies of the 1920s would bind evolutionism to theological liberalism in the minds of the fundamentalist faithful, but in Warfield’s lifetime, during the mature years of his career, many orthodox theologians accepted biological evolutionism to some degree. Warfield himself took a trajectory somewhat apart. He followed scientific developments with an interested eye and cheerfully welcomed new discoveries in archaeology, astronomy, and paleontology.17 But he retained a wariness about the metaphysical and philosophical assumptions that often attended scientific claims. When evolutionary theory went through a decades-long crisis around 1900, a period Peter Bowler

James McCosh, Christianity and Positivism (New York: Robert Carter and Bros., 1871); Warfield, “Personal Recollections,” 652. 14 George Macloskie, quoted in William Milligan Sloane, The Life of James McCosh (New York: Charles Scribner’s Sons, 1896), 124. 15 Charles Hodge, What Is Darwinism? (New York: Scribner, Armstrong & Co., 1874), 176-77. 16 See Gundlach, Process and Providence, chap. 4. 17 For example, B. B. Warfield, “Great Babylon the Mother of Us All,” in Selected Shorter Writings of Benjamin B. Warfield, ed. John M. Meeter (Nutley, NJ: Presbyterian & Reformed, 1970–1973), vol. 1. 13

B. B. Warfield (1851–1921)

65

calls the “eclipse of Darwinism,”18 Warfield gave particular attention to the question of biological evolution, for at that stage the science was focusing not on the factuality of transmutation but the how—offering many rival theories, none of which seemed to stick.19 Most were methodologically naturalistic. Some, like Henri Bergson’s, were not. McCosh, Charles Hodge, and many other conservative evangelicals had identified Darwin’s denial of supernatural agency as the crux of the evolution question theologically. Now when natural selection seemed an insufficient explanation (the neo-Darwinian synthesis of natural selection with Mendelian genetics would come only after 1930), yet liberal theologians offered a Christianity shorn of the miraculous, Warfield identified supernaturalism as the key issue of his day. A staunch confessional Calvinist, Warfield turned to Calvin and the Reformed scholastics for resources to explore the modes of God’s supernatural activity in this world. Framing the Evolution Question for Theological Students In his recollections of President McCosh, Warfield said it was at about the age of thirty that he himself “fell away” from McCosh’s “orthodoxy”— namely, that natural selection was God’s design for the origin of species.20 A zealous opponent of attempts to revise the Westminster Confession, Warfield here poked fun at the hidebound “orthodoxy” of scientific opinion when it came to evolution. It is important to note that Warfield was not denying transmutation; rather, he denied the sufficiency of natural selection, even if God-ordained, to achieve it. Peter J. Bowler, The Eclipse of Darwinism: Anti-Darwinian Evolution Theories in the Decades Around 1900 (Baltimore, MD: Johns Hopkins University Press, 1983). 19 Warfield wrote numerous reviews of theological books that treated the evolution question in light of contemporary scientific developments. His articles for The Bible Student show that he kept abreast of some of those scientific developments himself. These are all collected in Noll and Livingstone, Evolution, Science, and Scripture. 20 “No, he did not make me a Darwinian, as it was his pride to believe he ordinarily made his pupils. But that was doubtless because I was already a Darwinian of the purest water before I came into his hands, and knew my Origin of Species, and Animals and Plants under Domestication, almost from A to Izard. In later years I fell away from this, his orthodoxy. He was a little nettled about it and used to inform me with some vigor—I am speaking of a time some thirty years agone!—that all biologists under thirty years of age were Darwinians. I was never quite sure that he understood what I was driving at when I replied that I was the last man in the world to wonder at that, since I was about that old myself before I outgrew it.” Warfield, “Personal Recollections,” 652. 18

66

Bradley J. Gundlach

By the time Warfield took the chair of theology at Princeton at the age of thirty-six, he had concluded that “any form of evolution which rests ultimately on the Darwinian idea is very improbable as an account of how God has wrought in producing species.” The scientific arguments were far from convincing. He hastened to add, though, that he was not denying that species came by descent. Rather, if they did, it was not merely by the powers God had built into nature; that would be mere deism, and Christianity demanded something more. “A thoroughgoing evolutionism cannot be held in entire consistency with some other Christian doctrines,” especially “the doctrine of the substantiality and immateriality of the soul” and its life after the death of the body. The soul must have an origin above nature. Warfield declared himself “a pure agnostic” on the question of “whether species have in any way come by descent,” and he even stated that “the sole passage [in Gen 1–2] which appears to bar the way” to an evolutionary account of man “is the very detailed account of the creation of Eve.” Yet he concluded: The upshot of the whole matter is that there is no necessary antagonism of Christianity to evolution, provided that we do not hold to too extreme a form of evolution. To adopt any form that does not permit God freely to work apart from law and that does not allow miraculous intervention (in the giving of the soul, in creating Eve, etc.) will entail a great reconstruction of Christian doctrine, and a very great lowering of the detailed authority of the Bible. But if we condition the theory by allowing the constant oversight of God in the whole process, and his occasional supernatural interference for the production of new beginnings by an actual output of creative force, producing something new, i.e., something not included even in posse in preceding conditions, we may hold to the modified theory of evolution and be Christians in the ordinary orthodox sense. I say we may do this. Whether we ought to accept evolution, even in this modified sense, is another matter, and I leave it purposely an open question.21

Here, then, in 1888, though Warfield considered the transmutation theory far from proven, he allowed the possibility that Adam was the B. B. Warfield, MS lecture, “Anthropology,” systematic theology course, junior year (PTS Special Collections), published in part as “Evolution or Development,” in Noll and Livingstone, Evolution, Science, and Scripture, 114-31. Quotations are from p. 125.

21

B. B. Warfield (1851–1921)

67

progeny of brute ancestors—provided that God intruded direct, supernatural power in giving him a soul. It is striking that Warfield stated these views to first-year seminarians in the required theology sequence. But he also warned against “a thoroughgoing evolutionism” that would limit God’s activity in the world to indirect influence always and only through natural laws, and he countered the growing scientific claim that evolution had attained the status of proven fact. Turning to Current Science and the Reformed Theological Heritage Over the next decade it became increasingly clear to Warfield that more theological work needed to be done here. On the one hand, he sought input on the question of the soul’s origins from various contemporary disciplines, including biologists and philosophers; on the other hand, he found particularly useful the work of Reformed theologians of centuries past, Johannes Wollebius and John Calvin. Perhaps the most surprising episode in Warfield’s exploration of questions related to evolution came in 1890 when he was launching a new theological journal, the Presbyterian and Reformed Review (PRR), just as he was embarking on several years of intense controversy in the Presbyterian church over the higher criticism of the Bible, the revision of the Westminster Confession, and the heresy trials of Charles Briggs and Henry Preserved Smith.22 Warfield organized a symposium of several articles for the PRR around the theme “What Is Animal Life?” to explore the possibility that the human soul had arisen from, or at least had something in common with, an immaterial component in animals. Intent on guarding the duality of mind and body in human beings, Warfield invited input from science and philosophy on the possibility that the soul had evolved—not, indeed, from body but from an immaterial principle in animal precursors—and published a diversity of views, strikingly, in his conservative new journal dedicated to

For a classic account of that crucial moment in Presbyterian history, see Lefferts A. Loetscher, The Broadening Church: A Study of Theological Issues in the Presbyterian Church Since 1869 (Philadelphia: University of Pennsylvania Press, 1954), chaps. 5–8.

22

68

Bradley J. Gundlach

­Calvinist orthodoxy.23 Though he would ultimately reject the idea of a separate soul evolution, this episode illustrates Warfield’s openness to light from science and philosophy.24 By 1896, the forward momentum of what would come to be called theological liberalism was undeniable, and it was clear to Warfield that “the magic watchword of ‘evolution’” was aiding and abetting the rapid recasting of Christianity in nonmiraculous terms. He devoted his address at the opening of the seminary year to the importance of “Christian Supernaturalism” as the crying need of the hour in response to movements in both science and theology. How absolutely determinant the conception of evolution has become in the thinking of our age, there can be no need to remind ourselves. It may not be amiss, however, to recall the antisupernaturalistic root and the antisupernaturalistic effects of the dominance of this mode of conceiving things, and thus to identify in it the cause of the persistent antisupernaturalism that at present characterizes the world’s thought.25

Rather than ask, “What kind and measure of supernaturalism does the Christianity of Christ require?” Warfield lamented that too many people asked, “How little of the supernatural may be admitted and yet men continue to call themselves Christians?” They looked “to de-supernaturalize Christianity so as to bring it into accord with the prevailing world-view.” Hence the higher-critical view of the Bible as a record of the “naturalistic B. B. Warfield, “What Is Animal Life?,” Presbyterian and Reformed Review 1 (1890): 441-61; Gundlach, Process and Providence, 239-40. Contributors included paleontologists William Berryman Scott (Charles Hodge’s grandson and a non-Darwinian evolutionist professor at Princeton College) and John William Dawson (the last great North American scientist to hold out against evolution), theologians William G. T. Shedd and John DeWitt, philosopher John Dewey (early in his career), and theistic evolutionist Henry Calderwood. 24 The coup de grâce to the idea of soul evolution seems to have come from James Orr’s Stone Lectures at Princeton in 1903, a work the Princetonians cited repeatedly. Orr argued that the origin of man could not be divided into two separate works, the production of the body and the production of the soul, for the step from brute to man required physical changes adequate to support the new spiritual endowments, and the Bible treated humankind as a unity of soul and body, the separation of which (by the fall) was utterly unnatural. James Orr, God’s Image in Man, and Its Defacement in the Light of Modern Denials (New York: A. C. Armstrong, 1905). 25 B. B. Warfield, “Christian Supernaturalism,” in The Works of Benjamin B. Warfield, vol. 9, Studies in Theology, ed. Ethelbert D. Warfield et al. (New York: Oxford University Press, 1932; repr., Grand Rapids, MI: Baker Books, 1991), 28. The address first appeared in the Presbyterian and Reformed Review 8 (1897): 58-74.

23

B. B. Warfield (1851–1921)

69

development of the religion of Israel”; hence also the denial of the virgin birth, the bodily resurrection of Christ, and his miracles.26 Observe that Warfield pointed here to four of the future “five fundamentals.”27 But unlike many fundamentalists of the 1910s and 1920s, Warfield did not blame liberalism on the transmutation theory itself— rather, it was an antisupernaturalistic worldview that was the trouble. Indeed, even in this address Warfield went on pointedly to reaffirm “the reality and efficiency of second causes.” God created “real substance endowed with real powers” that “really act and really produce their effects.” “Just because he believes that the universe was well-made,” the Christian “believes that the forces with which it was endowed are competent for its ordinary government and he traces in their action the divine purpose unrolling its faultless scroll.” Note here that unrolling is the precise English equivalent of the Latin evolving: Warfield went on to say that the Christian is free to trace out God’s use of second causes “in all the products of time” and even to welcome the words of Tennyson: This solid earth whereon we tread, In tracts of fluent heat began, And grew to seeming random forms, The seeming prey of cyclic storms, Till at the last arose the man.28

Warfield was saying that to the extent nature “unrolled,” it was under the providential government of God, whose eternal decree the forces of nature serve. But “in our conception of a supernatural God, we must not erect His providential activity into an exclusive law of action for Him, and refuse to allow of any other mode of operation.” And here Warfield suggested a striking possibility: Who can say, for example, whether creation itself, in the purity and absoluteness of that conception, may not be progressive, and may not correlate Warfield, “Christian Supernaturalism,” 29-30. Of the famous “five fundamentals”—(1) the inerrancy of Scripture, (2) the virgin birth, (3) the substitutionary atonement, (4) the bodily resurrection, and (5) the biblical miracles—only (3) is absent here. The list first appeared fourteen years later, at the 1910 General Assembly of the PCUSA. 28 Warfield, “Christian Supernaturalism,” 36. The poem in question is Tennyson’s “In Memoriam,” famous for its image of “nature red in tooth and claw.” 26 27

70

Bradley J. Gundlach

itself with and follow the process of the providential development of the world, in the plan of such a God—so that the works of creation and providence may interlace through all time in the production of the completed universe?29

But within five years, Warfield came to view “creation by evolution” as a contradiction. Though creation and providence might “interlace” throughout the geological ages, the distinction between their modes was theologically important. Further reading persuaded him that creation and evolution were rightly understood as mutually exclusive terms. The theologian who so persuaded him was the early seventeenth-century Calvinist, Johannes Wolleb of Basel (1586–1629), better known by the Latinized name Wollebius. In his Compendium Theologiae Christianae (1626), Wollebius carefully distinguished between creation and mediate creation—and Warfield found in this distinction the key to understanding important differences among views simplistically lumped together as “theistic evolution.” He applied Wollebius’s conceptual scheme in a prolonged editorial note in The Bible Student, a new magazine intended for a lay audience. His use of that venue suggests that Warfield was eager to reach not just pastors and theologians but a religious public concerned to keep their faith in the Scriptures. In not too many years The Bible Student would evolve into a pretty strident fundamentalist publication. But in 1901 it carried one of Warfield’s most careful and significant writings on the evolution question. Some who called themselves theistic evolutionists, Warfield observed, were effectively deists: they pictured God as creating the original worldstuff and its laws, then letting them carry out the work of producing the habitable earth and its panoply of life forms over eons of time. To rule out the intervening hand of God into the natural world was to contradict the Christian gospel of supernatural salvation. Other theistic evolutionists, however, spoke of evolution as an instance of “mediate creation”—such as McCosh, the Duke of Argyll, and the Roman Catholic J. A. Zahm, whose book Evolution and Dogma Warfield made his prime example. They rightly wanted to preserve God’s prerogative to intrude supernatural Warfield, “Christian Supernaturalism,” 37.

29

B. B. Warfield (1851–1921)

71

power into natural history at such points as the creation of life and of consciousness, not to mention the biblical miracles and the miracle of spiritual rebirth. Warfield fully sympathized there, but on reading ­Wollebius he decided that it was a mistake to classify evolution as a kind of mediate creation. Here is his discussion: “Creation,” he [Wollebius] says, “is that act by which God, for the manifestation of the glory of His power, wisdom, and goodness, has produced the world and all that is in it,”—we relapse now into his Latin—“partim ex nihilo, partim ex materia naturaliter inhabili,”—that is to say, in part, out of nothing, and in part out of preexisting material indeed, but material not itself capable of producing this effect. Again: “to create is not only to make something out of nothing but also ex materia inhabili supra naturae vires aliquid producere,” to produce something out of this inapt material, above what the powers intrinsic in it are capable of producing.30

Mediate creation was God’s act of intruding supernatural power into the normal web of natural causes that God, by his providence, was continually upholding and governing. When Jesus made water into wine, that was a mediate creation: throughout the event the jars contained the liquid, but the liquid changed from water to wine in a way water itself is never capable of doing. The water was not the means of the miracle but rather the object acted on by miraculous power. It was there by God’s providence, and it became wine not by creation out of nothing but by an act of mediate creation. Those three modes of the supernatural—­ providence, creation, and mediate creation—needed to be distinguished. Each was supernatural: providence indirectly, as God acts through created means operating according to their created natures; creation ex nihilo utterly directly; and mediate creation directly but making use of something already created, changing it in a way that it was not, as created, capable. Most miracles were acts of mediate creation, and in these days of “curbing the supernatural” it was of paramount importance to preserve the supernaturalness of miracles and not to fancy it a small thing to recast them as mere providences. Especially in the salvation of the individual sinner, it was crucial to recognize our need of a B. B. Warfield, “Editorial Notes,” Bible Student, n. s., 4, no. 1 (July 1901): 5.

30

72

Bradley J. Gundlach

supernatural regeneration rather than to tell us to imitate Christ by natural human effort. Warfield would find this notion of mediate creation useful in many ways, including his argument for the role of humanity’s common reason in Christian apologetics, in contrast to the views of Abraham Kuyper. It informed his doctrine of the role of human effort in sanctification versus the perfectionism of the “Victorious Life” movement. It elaborated his long-standing doctrine of God’s activity in the mysterious process of inspiration, concursively producing an inerrant Bible through the agency of imperfect human authors, with the intrusion of supernatural power from above.31 When Warfield turned to consider John Calvin’s doctrine of creation in 1915, he found that the great reformer made a point of reserving creation ex nihilo for the original world-stuff and the souls of humankind. All else Genesis treats as forming—“gradual modelling into form,” to use Warfield’s interpolation—rather than creating as such. In answer to the scoffer’s charge that an omnipotent God should not have needed to take six days to make the world, Calvin calls his readers’ minds, in Warfield’s words, To dwell on the condescension of God in distributing His work into six days so that our finite intelligence might not be overwhelmed with its contemplation; and on the goodness of God in thus leading our thoughts to the consideration of the rest of the seventh day; and above all on the paternal care of God in so ordering the work of bringing the world into being as to prepare it for man before He introduced him into it.32

Thus Calvin teaches “that God perfected the world by process (progressus, I.xiv.2) . . . not for His own sake, but for ours.” Wanting us to be able to contemplate his goodness in creation, God protracted the process so that our minds could perceive his sequence of loving acts in preparing our home. Calvin is not interested in the details of how God created the world, except as to encourage loving contemplation of these themes. Yet Warfield B. B. Warfield, “Introductory Note” to Francis Beattie’s Apologetics: Or the Rational Vindication of Christianity (Richmond, VA: Presbyterian Committee on Publications, 1903), 19-32; B. B. Warfield, “The Victorious Life,” in Works, vol. 8, Perfectionism, Volume 2, 561-611; Jeffrey A. Stivason, From Inscrutability to Concursus: Benjamin B. Warfield’s Theological Construction of Revelation’s Mode from 1880 to 1915 (Phillipsburg, NJ: P&R Publishing, 2017). 32 B. B. Warfield, “Calvin’s Doctrine of the Creation,” in Works, vol. 5, Calvin and Calvinism, 298. 31

B. B. Warfield (1851–1921)

73

finds in Calvin’s exposition a surprising thing: in reserving the concept of creation proper for the original world-stuff and human souls alone, Calvin teaches that “the world as we now see it . . . has been evoked by the progressive acts of God.” Those progressive acts of God, Warfield then observes, Calvin does not refer to mediate creation as Wollebius and other Reformed divines would do a few generations later. Calvin expressly refuses the idea of subsequent infusions of creative energy from outside God’s providential activity; the fashioning of the world and of life forms after the initial creation ex nihilo was entirely providential, up until the creation of Adam’s soul.33 And so Warfield makes his startling claim: It should scarcely be passed without remark that Calvin’s doctrine of creation is, if we have understood it aright, for all except the souls of men, an evolutionary one. The “indigested mass,” including the “promise and potency” of all that was yet to be, was called into being by the simple fiat of God. But all that has come into being since—except the souls of men alone—has arisen as a modification of this original world-stuff by means of the interaction of its intrinsic forces. . . . To him God is the prima causa omnium and that not merely in the sense that all things ultimately—in the world-stuff—owe their existence to God; but in the sense that all the modifications of the world-stuff have taken place under the directly upholding and governing hand of God, and find their account ultimately in His will. But they find their account proximately in “second causes”; and this is not only evolutionism but pure evolutionism.34

What the scientist called evolution, often imagining that evolution somehow took the place of God’s activity, Warfield called providence— and not the providence of the deist, who thought of natural law as providence’s only mode. With Calvin, Warfield held a very high doctrine of providence as the constant forth-putting of divine power in upholding the creatures in all their created powers. It is a mistake to cite this exposition of Calvin as proof that Warfield was placing the imprimatur of Calvin on the scientific theory of biological or cosmological evolution. But he was indeed finding theological precedent for a Christian and a Calvinist to allow evolution, up to but not including the human soul. Warfield, “Calvin’s Doctrine of the Creation,” 298-300. John Murray argues that Warfield has misread Calvin; see note 54 below. 34 Warfield, “Calvin’s Doctrine of the Creation,” 304-5. 33

74

Bradley J. Gundlach

But more needs to be said. Warfield’s exposition of Calvin on the lower creation—the world and its life forms—occupied about nine pages of the article. He spent three times that space, twenty-seven pages, expounding Calvin’s doctrine of the angels and the use God made of them as “second causes” in forming things as they now are. After all, what would the evolution of the world and life in just six days look like? Not the immensely protracted, gradual, and this-worldly natural process envisioned by the scientists. Warfield was an old-earther, but Calvin was not—so Calvin’s doctrine of creation, providential in its mode, envisioned a very different kind of providence than we would expect: an angelic one. “God executes His works of providence through the intermediation of second causes; for this is the very definition of providence.” But “among these second causes there are always personal as well as impersonal agencies,” and Calvin conceives “that all the works of God’s providence are wrought through the intermediation of angels.”35 Warfield himself did not credit angels with causing or guiding evolution. But this aspect of the article, passed over by commentators on Warfield’s view of evolution,36 makes it clear that in expounding Calvin, Warfield was not recommending what we would normally consider an evolutionary view of earth history. He takes pains to emphasize “the vividness of his [Calvin’s] sense of the spiritual environment in which our life is cast.” We see here that he conceived the universe as in all its operations moving on under the guiding hand of these superhuman intelligences. This is as much as to say that there was no dualism in his conception of the universe: he did not set the spiritual and physical worlds, or the earthly and supramundane worlds, over against one another as separate and unrelated entities. He conceived them as all working together in one unitary system, acting and interacting on one another. And he accustomed himself to perceive beneath the events of human history—whether corporate or individual—and beneath the very operations of physical nature—not merely the hand of God, Warfield, “Calvin’s Doctrine of the Creation,” 322. Noll and Livingstone’s anthology does not reproduce this section on the angels. They do attempt a one-paragraph summary of it but deemphasize (at best) this stress on angelic instrumentality in the formation and maintenance of the physical universe. B. B. Warfield: Evolution, Science, and Scripture, 311. Even Zaspel, in his brief treatment of Warfield’s article in Theology of B. B. Warfield (384), overlooks entirely the emphasis Warfield here gives on Calvin’s appeal to the agency of angels.

35 36

B. B. Warfield (1851–1921)

75

upholding and governing; but the activities of those “hands of God” who hearken to His voice and fulfil His word, and whom He not only charges with the care of His “little ones,” and the direction of the movements of the peoples, but makes even “winds” and a “flaming fire.”37

In providence “God thus universally operates through the instrumentality of subordinate intelligences.” God could have acted immediately, but he chose to operate through the angels in order “to give us not only His protection but the sense of His protection. Dealing with us as we are, not as we ought to be, He is willing to appeal to our imagination and to comfort us in our feeling of danger or despair by enabling us to apprehend, in our own way, the presence of His grace.” He promises not only that he will care for us but that he sends (now quoting Calvin) “‘innumerable escorts to whom He has given charge to secure our safety (§ 11).’”38 God wants us to contemplate the host of angelic intermediaries whom he sends not only occasionally, but constantly, such that the very world we inhabit was and is framed and formed by these guardians of our souls doing God’s bidding. In light of this striking re-enchantment of the universe, Warfield hardly appears as the rationalist that he is sometimes accused of being. If he is endorsing here Calvin’s providentialism as evolutionism, he is also endorsing a vivified spiritual universe of providential evolution by angelic beings. No wonder his later reviews of books on the evolution question display an annoyance at the antisupernaturalistic bias of scientists who have no viable theory of the “origin of the fittest” yet insist that any explanation of such origin must avoid recourse to the supernatural. It is there that Fred Zaspel finds much supporting evidence for his contention that the progress of Warfield’s thought was away from, not toward, support for evolution.39 Warfield, “Calvin’s Doctrine of the Creation,” 323. Warfield, “Calvin’s Doctrine of the Creation,” 324, quoting Calvin’s Institutes, 1.14.11. 39 B. B. Warfield, review of Darwinism To-day, by Vernon L. Kellogg, Princeton Theological Review 6 (1908): 647; Gundlach, Process and Providence, 287-92. While Noll and Livingstone (Evolution, Science, and Scripture, 36) state that Warfield’s evolutionism grew surer in the 1900s, Zaspel (Theology of B. B. Warfield, 377-81) argues the opposite. I too find Warfield’s exasperation with scientific dogmatism to grow after 1900 and his confidence in any given evolutionary explanation to dwindle. But at this time the “eclipse of Darwinism” was at its fullest, after all—yet Warfield never came out against the theory of descent, just its abuses. 37 38

76

Bradley J. Gundlach

But Warfield never repudiated the doctrine of evolution. He never endorsed it outright, either. He allowed it, and in my opinion expected the transmutation of species to be proven eventually, for he saw signs everywhere that God’s standard operating procedure was gradual and progressive. This may be seen in the work of colleagues at Princeton Seminary. Geerhardus Vos, Princeton’s first professor of biblical theology, taught a strikingly developmental view of the history of revelation in Scripture. In his inaugural address he declared, “God has not communicated to us the knowledge of the truth as it appears in the calm light of eternity to His own timeless vision. . . . The self-revelation of God is a work covering ages. . . . The truth comes in the form of growing truth, not truth at rest.” The task of biblical theology is “the exhibition of the organic progress of supernatural revelation in its historic continuity and multiformity.” Vos employed an analogy of growth from seed to full flower, in which every stage is perfect though completion comes only at the end. This was the only way to reconcile the two great facts of Scripture, divine perfection and historical process.40 Another example is George T. Purves, professor of New Testament, who urged readers to recognize “the evolution of revelation” in the course of biblical history, “as God wrought through human media the perfect disclosure of saving truth.”41 It is worth noting that Warfield enjoyed close personal friendships with both of these men and took the lead in bringing them to join him on the seminary faculty. Warfield himself employed this analogy of organic growth to the progress of theology. Theology was an “organic growth,” “the ripened fruit of the ages.”42 The Westminster Confession, the latest of the classic Protestant creedal statements, far from suffering from its distance in time from the original impulse of the Reformation, was the beneficiary of a Geerhardus Vos, “The Idea of Biblical Theology as a Science and as a Theological Discipline” (1849), in Redemptive History and Biblical Interpretation: The Shorter Writings of Geerhardus Vos, ed. Richard B. Gaffin (Phillipsburg, NJ: Presbyterian and Reformed, 1980), 7, 10-11, 13-14. See my treatment in Process and Providence, 209-13. 41 George Tybout Purves, review of Messianic Prophecy, by Edward Riehn, Presbyterian and Reformed Review 3 (1892): 554. In rather Warfieldian fashion, Purves embraced what he could of biblical criticism’s historical point of view and interest in finding “the supernatural in the natural” while yet maintaining the verbal inspiration of the text. 42 B. B. Warfield, “The Idea of Systematic Theology,” in Works, vol. 9, Studies in Theology, 75-77. 40

B. B. Warfield (1851–1921)

77

long process of careful exploration of the ramifications of Reformation principles. It represented “the ripest fruit of Reformed creed-making.”43 And perhaps the best example of Warfield’s pervasive developmentalism is found in his response to the perfectionism of “let go and let God” movements in his day: Men are unable to understand why time should be consumed in divine works. Why should the almighty Maker of the heaven and earth take millions of years to create the world? Why should He bring the human race into being by a method which leaves it ever incomplete? Above all, in his recreation of a lost race, why should He proceed by process? Men are unwilling that either the world or they themselves should be saved by God’s secular methods. They demand immediate, tangible results. They ask, Where is the promise of His coming? They ask to be themselves made glorified saints in the twinkling of an eye. God’s ways are not their ways, and it is a great trial to them that God will not walk in their ways. They love the storm and the earthquake and the fire. They cannot see the divine in “a sound of gentle stillness,” and adjust themselves with difficulty to the lengthening perspective of God’s gracious working. They look every day for the cataclysm in which alone they can recognize God’s salvation.44

In nature and in grace, God proceeds by process. Warfield was careful to guard the supernaturalism so crucial to a Christian view of the world and so crucial to the gospel of Christ that offered spiritual rebirth, a new creation. But the nature of supernatural agencies was somewhat scrutable, and theologians past could help us understand “God’s secular methods” as something far more spiritual than theistic evolutionists typically teach. Theistic Evolution and Human Origins By now it should be clear that Warfield hardly qualifies as a “theistic evolutionist,” even in the usual broad and imprecise sense of one who affirms the accepted scientific picture with an added veneer of belief that “God did it.” Thus far Zaspel is correct—Warfield never avowed B. B. Warfield, “The Westminster Assembly and Its Work,” in Works, vol. 6, The Westminster Assembly and Its Work, 58-59. 44 Warfield, “The Victorious Life,” in Works, vol. 8, Perfectionism, Volume 2, 561. 43

78

Bradley J. Gundlach

any kind of transmutationism as proven fact, and indeed on the topic of human origins (the only really theologically significant application of the theory, in Warfield’s view) he insisted on a historical Adam and Eve, progenitors of the entire human race, endowed originally with righteousness, whose fall ruined all humankind. However, Warfield’s position on both theistic evolution and human origins was more complex and theologically specific than that, and the particulars are worth exploring briefly here. Warfield first addressed the evolution question in 1888, in lectures to entering theology students on the doctrine of man, quoted at length above. He used them with little change for the next three decades, until he handed over the class to his assistant, Caspar Wistar Hodge Jr., in 1906/1907. Historians have used these notes to prove both Warfield’s openness to evolution and his skepticism toward it. In them Warfield clearly allows a limited evolutionism, limited by God’s superintendence and “occasional supernatural interference.” He even allows the possibility of an evolutionary element in the creation of Adam’s body, though not his soul. Yet he insists that full-blown evolutionism’s elimination of the miraculous cannot comport with orthodox Christianity, and it does violence to the Bible. Just on this point Warfield is careful to distinguish merely theistic evolution from a more acceptable version. Admitting that an alwaysand-only-natural evolution presents no conflict with theism—mere belief in “an everywhere present and active God who nevertheless acts only according to law”—Warfield continues, “But to be a theist and a Christian are different things. And a thoroughgoing evolutionism cannot be held in entire consistency with some other Christian doctrines.” 45 Surely this is a solid disavowal of “theistic evolution,” properly so-called. Warfield drew out this distinction more fully in “The Antiquity and Unity of the Human Race.” In this article, published in 1911, he reaffirms the conclusion of his Old Testament mentor William Henry Green that the genealogies of the antediluvians are not intended for use in dating the creation of humankind, and indeed that the antiquity of the human Warfield, “Victorious Life,” 125.

45

B. B. Warfield (1851–1921)

79

race is of no theological concern.46 This would of course be useful in accommodating the long ages required by evolutionary theory—yet clearly this is not Warfield’s sole purpose,47 for he goes on to argue that the Bible records the extreme longevity of the antediluvians in order to impress us with their great vitality, due presumably to their genetic nearness to the original unfallen state of humanity.48 But the burden of the article is to show that while the antiquity of the human race is of no theological concern, the unity of the race in Adam, its progenitor, most emphatically is. The unity of the race in Adam is “the postulate of the entire body of the Bible’s teaching—of its doctrine of Sin and Redemption alike: so that the whole structure of the Bible’s teaching, including all that we know as its doctrine of salvation, rests on it and implicates it.” Warfield insists that the Bible teaches that Adam and Eve, as well as humanity’s “second father, Noah,” were real people “from whose fruitfulness and multiplication all the earth has been replenished.” Scripture treats all humankind “as, from the divine point of view, a unit,” sharing “not only in a common nature but in a common sinfulness, not only in a common need but in a common redemption” through Christ, the second Adam.49 Any Christian version of evolution must affirm a On Green’s argument and its wide influence, see Ronald L. Numbers, “‘The Most Important Biblical Discovery of Our Time’: William Henry Green and the Demise of Ussher’s Chronology,” Church History 69 (2000): 257-76. 47 Warfield states that the gaps in the genealogies could accommodate even two hundred thousand years between Adam and Abraham, so little do those genealogies have to do with chronology. B. B. Warfield, “Evolution or Development,” in Noll and Livingstone, Evolution, Science, and Scripture, 130-31. Original in his MS notes on Anthropology, Special Collections, Princeton Theological Seminary Library. But after all, that would not speak to the question of evolution up to the time of humans, since Adam was the first human. Warfield considers the Bible to be silent on the duration of earth history before Adam. 48 B. B. Warfield, “The Antiquity and Unity of the Human Race,” in Works, vol. 9, Studies in Theology, 241: “When we are told of any man that he was a hundred and thirty years old when he begat his heir, and lived after that eight hundred years begetting sons and daughters, dying only at the age of nine hundred and thirty years, all these items cooperate to make a vivid impression upon us of the vigor and grandeur of humanity in those old days of the world’s prime.” Here Warfield is repeating an argument he made in a more popular magazine, The Bible Student, in 1903. See Noll and Livingstone, Evolution, Science, and Scripture, 220-21. 49 Warfield, “Antiquity and Unity of the Human Race,” 251-52. Cf. 258: “So far is it from being of no concern to theology, therefore, that it would be truer to say that the whole doctrinal structure of the Bible account of redemption is founded on its assumption that the race of man is one organic whole, and may be dealt with as such. It is because all are one in Adam that in the matter of sin there is no difference, but all have fallen short of the glory of God (Rom. iii. 22 f.), and 46

80

Bradley J. Gundlach

real, historical Adam, just as Christianity depends on a real, historical Redeemer, Jesus Christ. And thankfully, Warfield points out, evolutionary theory can serve Christianity here, for it has put to rest all the racist theories of polygenism, preadamism, and coadamism that flourished in the nineteenth century.50 What, then, would the properly Christian evolutionism that Warfield allowed, as compared to a merely theistic evolutionism, look like—­ particularly in the case of humankind? Warfield never affirmed this ­position outright and never sketched it even in outline. Our best hope to find it lies in the unpublished anthropology lectures of his successor, Caspar Wistar Hodge Jr., begun in 1906/1907 and enlarged thereafter. Hodge discusses at length the exegetical, methodological, and worldview issues involved in the mode of man’s origin, building his case especially around the distinctions Warfield drew among creation, mediate creation, and providence (evolution), and concludes as follows. (I both quote and paraphrase here, enumerating for clarity.) 1. Adam’s body may be the product of mediate creation—from a clod of earth. 2. Or it may be the product of mediate creation from a preceding animal, but that animal “did not contain the potency of developing into the human body simply under God’s providential control of the second causes.” 3. “Or again if man’s body were evolved from lower organisms, then it was not created in any sense, but was simply the result of second causes under Divine control. But then we should have to admit the creative or supernatural intrusion in the case of the first man’s soul, as well that in the new man there cannot be Greek and Jew, circumcision and uncircumcision, barbarian, Scythian, bondman, freeman; but Christ is all and in all (Col. iii. 11). The unity of the old man in Adam is the postulate of the unity of the new man in Christ.” It is worth adding here that Warfield is not addressing the question of human evolution directly but only as it impacts the question of the unity of the race. 50 Warfield, “Antiquity and Unity of the Human Race,” 252-56. Warfield showed considerable familiarity with those theories and their pedigree going back to classical times. For details of these theories and their tortuous interrelationship with apologetics and racism, see David N. Livingstone, Adam’s Ancestors: Race, Religion, and the Politics of Human Origins (Baltimore, MD: Johns Hopkins University Press, 2008).

B. B. Warfield (1851–1921)

81

as the Scripture in Gen 2, we have seen, makes man as a whole due to God’s creative and not merely controlling, activity.” 4. Adam’s soul may have been an immediate creation—out of nothing. 5. Or it may have involved a previous form of soul life—but in that case it must have been a mediate creation, not an evolution. “If there be a genetic connection with lower forms of soul life, those forms do not contain the power or potency to produce it simply under God’s guidance. But while according to this hypothesis the soul would join on to the series of lower forms of soul life, yet God’s creative activity would intrude into the series of second causes and as it were make supernaturally or creatively something new, though making it out of a preceding form of life—i.e., a mediate creation analogous to the turning of water into wine.” 6. “We think that the Scripture account leaves room for these alternatives.”51 7. Thus, as Warfield’s assistant and colleague in the theology department for the last fifteen years of Warfield’s life, Hodge taught that Scripture allowed the possibility of prehuman ancestry for Adam’s body, whether by outright evolution (natural causes under God’s providential control) or mediate creation (intrusion of supernatural power to change the brute body into Adam’s body). He also taught, however, that the Bible ruled out the possibility that Adam’s soul came simply by evolution. It must have come by God’s creative act, though that might still be a mediate creation, involving God’s creative power acting on a “lower form of soul life.” It is hard to imagine Warfield allowing this teaching if he did not approve of it. Certainly, Hodge gave every indication of believing himself to be continuing Warfield’s views.

Caspar Wistar Hodge Jr., MS Notes on Anthropology (systematic theology course, middle year), Lecture I: “The Mode of Man’s Origin,” C. W. Hodge Jr. Papers, Box 1 (Special Collections, Princeton Theological Seminary Library). Elsewhere in these lectures Hodge explicitly rules out “theistic evolution” as an option for the Christian, since that category envisions a natural history with no breaks, no intrusion of supernatural creative power ever, as exemplified in the work of Otto Pfleiderer.

51

82

Bradley J. Gundlach

Conclusion Warfield’s position on evolution and particularly on human origins defies easy categorization, especially in terms of the labels available today. He certainly denied materialistic and even theistic evolution since those theories would rule out the existence of God or his power to intrude supernaturally in nature and history. He attempted to set acceptable theological boundaries rather than commit himself to any given scientific theory, especially in view of the unsettled state of evolutionary science at the time—a wise policy worth following in any era. He brought spiritual beings (angels) into the picture of God’s work in creating and governing the developments of this world, something commentators on Warfield’s views have overlooked. And he enlisted the aid of theologians past, especially Wollebius, to clarify the modes of God’s supernatural activity as they applied to the evolution question. Warfield’s use of the categories of creation, mediate creation, and providence is certainly not the last word on the subject. In fact, John Murray, professor at Westminster Theological Seminary in the mid-twentieth century, disagreed with Warfield’s reading of Calvin on this very issue of mediate creation versus providence. Readers interested in mining the riches of older theologies for help with the evolution question should begin with Murray’s very helpful article.52 Warfield’s openness on the question of human brute ancestry and prehuman evolution is remarkable in view of his simultaneous recognition that evolutionary ideas underlay the New Theology and the theological liberalism that was about to embroil the church in the fundamentalist controversies. Yet this openness had definite, carefully chosen bounds, and we do well to imitate him in this regard as we face the question of human origins as Christians today. The scientific situation has changed, but the theological issues, if they have been rightly identified, have not. John Murray, “Calvin’s Doctrine of Creation,” Westminster Theological Journal 17 (1954): 21-43. Murray uses more of Calvin’s works than Warfield does, showing pretty conclusively that Warfield had Calvin wrong. He also gives a handy survey of the prevalence of the distinction between creatio ex nihilo and creatio ex materia in the seventeenth century. Not only Calvinists but Lutherans and Roman Catholics made use of the distinction. Intriguingly, Murray notes that Charles Hodge deployed the idea of mediate creation in a way that departed, probably unwittingly, from Turretin and the other Reformed scholastics (28n7).

52

B. B. Warfield (1851–1921)

83

Recommended Reading Primary Works Warfield, B. B. Evolution, Science, and Scripture: Selected Writings. Edited by Mark A. Noll and David N. Livingstone. Grand Rapids, MI: Baker Books, 2000. ———. Selected Shorter Writings. 2 vols. Edited by John E. Meeter. Phillipsburg, NJ: Presbyterian & Reformed, 1970/1973. ———. The Works of Benjamin B. Warfield. 10 vols. Edited by Ethelbert D. Warfield, William Park Armstrong, and Caspar Wistar Hodge. New York: Oxford University Press, 1927–1932. Reprint, Grand Rapids, MI: Baker Books, 1991.

Secondary Works Gundlach, Bradley J. Process and Providence: The Evolution Question at Princeton, 1845– 1929. Grand Rapids, MI: Eerdmans, 2013. Johnson, Gary L. W., ed. B. B. Warfield: Essays on His Life and Thought. Phillipsburg, NJ: P&R Publishing, 2007. Livingstone, David N. Darwin’s Forgotten Defenders: The Encounter Between Evangelical Theology and Evolutionary Thought. Grand Rapids, MI: Eerdmans, 1987. ———. Dealing with Darwin: Place, Politics, and Rhetoric in Religious Engagements with Evolution. The Gifford Lectures, 2014. Baltimore, MD: Johns Hopkins University Press, 2014. Noll, Mark A. “A Case Study: B. B. Warfield, Concursus, and Evolution.” In Jesus Christ and the Life of the Mind. Edited by Mark Noll. Grand Rapids, MI: Eerdmans, 2011. ———. The Princeton Theology: Scripture, Science, and Theological Method from Archibald Alexander to Benjamin Warfield. Grand Rapids, MI: Baker Academic, 1983. ———. “Reprise: B. B. Warfield and Concursus.” In Jesus Christ and the Life of the Mind. Edited by Mark Noll. Grand Rapids, MI: Eerdmans, 2011. Riddlebarger, Kim. The Lion of Princeton: B. B. Warfield as Apologist and Theologian. Bellingham, WA: Lexham Press, 2015. Zaspel, Fred. The Theology of B. B. Warfield: A Systematic Summary. Wheaton, IL: Crossway, 2010.

4 RU D O L F BU LT M A N N ( 1 8 8 4 – 19 76 ) Myth, Science, and Hermeneutics Joshua W. Jipp Rudolf Bultmann was a Lutheran theologian and New Testament scholar who completed his dissertation at the University of Marburg under Johannes Weiss. Bultmann would return to teach at the University of Marburg from 1921 until his retirement in 1951, where he would also develop a friendship with Martin Heidegger. Bultmann was also a member of the Confessing Church, a Protestant movement that opposed the rise of Nazism in Germany. Bultmann was a pioneer of the dialectical theology movement and is best known for his demythologization project in hermeneutics and biblical interpretation, although the precise relationship between these two programs remains a subject of scholarly debate.

A

s a pioneer of form criticism of the Gospels, popularizer of the   results of the history of religions school, interpreter of Paul’s theology, and writer of the most significant twentieth-century New Testament theology and commentary on the Fourth Gospel, not to mention a variety of explicitly theological contributions, there is no doubt that Rudolf Bultmann casts a long shadow on the field of New Testament studies. Bultmann’s more than thirty years of teaching and training students at Marburg (1921–1951) resulted in dozens of his students occupying major university positions in biblical and theological studies both in Germany and in North America, including Ernst Käsemann, Günther Bornkamm, Hans Conzelmann, and Helmut Koester. It is difficult to disagree with William Baird’s assessment that Rudolf Bultmann is “the most important NT scholar of the twentieth century.”1 William Baird, History of New Testament Research, vol. 3, From C. H. Dodd to Hans Dieter Betz (Minneapolis: Fortress Press, 2013), 85.

1

Rudolf Bultmann (1884–1976)

85

Born in Wiefelstede, Germany (1884), Rudolf Bultmann had a happy childhood. While his mother was committed to historic Christian orthodoxy, his father, a Lutheran pastor, adopted numerous liberal positions that led to his alienation from many friends and family. Even as a young child, Bultmann “likewise freed himself from a number of old notions,” including “the orthodox concept of faith mediated by his religious education, which held that any religion must necessarily contain objective content.”2 Thus, very early on as a young theological student, unwilling to abandon Christianity and faith, Bultmann committed himself to re-envisioning theology and freeing it from its captivity to traditional doctrinal teachings on the Trinity, atonement, and miracles. Bultmann completed his studies at Marburg, but he also spent time as a student at Tübingen and Berlin and was thereby exposed to the teachings of Wilhelm Hermann, Adolf von Harnack, Hermann Gunkel, Wilhelm Heitmüller, Adolf Jülicher, and Johannes Weiss. After short teaching posts at the universities of Breslau and Giesen (1916–1921), Bultmann spent the majority of his academic career at Marburg (1921–1951), where he had written his dissertation on the influence of Cynic and Stoic philosophy on Paul’s letter to the Romans. While his later friendship with existential philosopher Martin Heidegger would play an important role in the development of his theology, very early in his academic career (ca. 1913) Bultmann had already jettisoned the task of dogmatics as one of reconstructing objective historical events and facts or theological-­ doctrinal concepts and had committed himself to “an existential participation” in “the eternal forces that reveal themselves in the temporal events of history.”3 While not given to engaging political and social events through his academic lectures and teachings, Bultmann belonged to the anti-Nazi Confessing Church.4 Bultmann’s theological genius, in part, is seen in his ability to set forth a creative theological reading of the Christian Scriptures, a reading that acted as a bridge between history and See further the remarkable biography by Konrad Hamman, Rudolf Bultmann: A Biography, trans. Philip E. Devenish (Salem, OR: Polebridge Press, 2013), 11. My biographical sketch here is largely dependent on Hamman’s remarkable biography. 3 That this is the case with respect to Bultmann’s early scholarship is shown beyond doubt by Hamman, Rudolf Bultmann, 55-63. 4 Baird, History of New Testament Research, 86. 2

86

Joshua W. Jipp

philosophy as well as exegesis and theology, through synthesizing aspects of the history of religions approach to Christian origins; existential philosophy; the dialectical theology of Barth, Gogarten, and Brunner; a particular Lutheran reading of Paul (indebted to Wilhelm Hermann); and Protestant liberalism.5 As we will see, Bultmann’s theological reading of the Scriptures often takes as its point of departure the relationship between mythical thinking and scientific thinking. Whereas myth and science are similar in that they both engage in objectifying thinking (more on this soon), modern science has, for Bultmann, rendered impossible a belief in the literal mythological statements of Scripture.6 Rudolf Bultmann: Myth, Science, and Hermeneutics Is it truly the case that scientific findings problematized the affirmation of the basic Christian doctrines of God as Creator—for example, God as creator of the cosmos, humanity and the world as in some way corrupted through the entrance of sin and evil, and ontological claims about humanity as created in God’s image? Rudolf Bultmann (1884–1976) certainly believed so. Bultmann provides an interesting and illuminating, albeit, in my view, problematic, way to think about how we might respond to these questions. Interesting and illuminating because Bultmann poses as sharply as anyone I know the challenges of the relationships between science and theology, faith and myth/worldview thinking, and subject and object. Problematic, however, because at the heart of Bultmann’s hermeneutics and theology are profound and untenable dualisms. In order to understand Bultmann’s attempt to affirm faith in God as Creator, we must set out his hermeneutical program of demythologizing and his understanding of science and myth. At the heart of Rudolf Bultmann’s understanding of the Christian faith and its continuing relevance in a scientific age is the appropriate interpretation of myth. For Bultmann, myth is a certain form of “thinking and speaking that objectifies the unworldly [Unweltliche] as something worldly [Welthafte].”7 Should one, On Bultmann’s influences, see Baird, History of New Testament Research, 86-93. See here also, David W. Congdon, The Mission of Demythologizing: Rudolf Bultmann’s Dialectical Theology (Minneapolis: Fortress Press, 2015), 662-63. 7 Rudolf Bultmann, “On the Concept of Myth,” in Congdon, Mission of Demythologizing, 853. 5 6

Rudolf Bultmann (1884–1976)

87

for example, portray the creation and origins of the world analogously to an emergence of a piece of art, one is using mythical language or human analogy to portray the divine. Bultmann moves from this claim to the belief that myth thereby does not intend to say anything about the world and its workings; rather, for Bultmann, myth is a primitive sort of objectifying thinking that is ultimately concerned with the individual human’s relationship to the cosmos. Mythical language about the world is indeed rendered obsolete if it is taken as a form of primitive science (i.e., as an early but crude version of natural scientific investigation), but ultimately myth, in Bultmann’s view, has as its purpose a form of speaking about God from within the realm of human existence. In other words, we are confronted here by a stark dualism between human existence and the physical-substantive cosmos in Bultmann’s thinking. Myth and science are not concerned with the same tasks or motivations, given that the former is ultimately concerned with the grounds of human existence.8 They do not seem to overlap in any way for Bultmann. “Guided originally by the question of one’s own existence, mythical thinking does not inquire, like distance-establishing scientific thinking, after the new, the interesting, and the strange in order to reduce them to the known and familiar,” says Bultmann, “but rather it inquires after the uncanny and the frightening . . . in order to secure itself against them.”9 Scientific thinking speaks of the world rationally, objectively, and from a distance (as opposed to myth, which, although it also engages in objectifying language, speaks of ultimate reality from within the ground of human existence). Operant within science, of course, is a divide between the human subject and the object; there is no concern for an existential encounter or a search for the truth of human existence in scientific thinking. In contrast to science, myth is ultimately concerned with human existence in its descriptions of the cosmos. Thus, Bultmann argues that myth is to be interpreted not “in cosmological terms but in anthropological

There is, however, some similarity between science and myth in that they both engage in objectifying thinking, although it is only the latter that attempts to speak about the truth of human existence. See here David W. Congdon, “Demystifying the Program of Demythologizing: Rudolf Bultmann’s Theological Hermeneutics,” Harvard Theological Review 110 (2017): 6-7. 9 Bultmann, “On the Concept of Myth,” 854. 8

88

Joshua W. Jipp

terms—or, better, in existentialist terms.”10 Yet for Bultmann myth still remains the vehicle for divine revelation since myth is ultimately concerned with the interpersonal, concrete, subjective encounter of God.11 Bultmann suggests that this is good news for those who would wish to remain Christians, as accepting myth on its own terms, or “repristinating the mythical world picture,” is simply impossible given that “all of our thinking is irrevocably formed by science” and would involve a “forced sacrificium intellectus.”12 Science and philosophy, for Bultmann, have demonstrated the impossibility of accepting myth as a primitive but nevertheless objectively truthful account of the world. One thinks of Bultmann’s now cliché but memorable way of saying this: “We cannot use electric lights and radios and, in the event of illness, avail ourselves of modern medical and clinical means and at the same time believe in the spirit and wonder world of the New Testament.”13 Bultmann is nothing if he is not clear about the impossibility of believing in miracles or wonders. Why have they become impossible? Modern science, according to Bultmann, says so. “The idea of wonder as miracle has become almost impossible for us today because we understand the processes of nature as governed by law. Wonder, as miracle, is therefore a violation of the conformity to law which governs all nature, and for us today this idea is no longer tenable.”14 Thus, the modern interpreter’s hermeneutical program is not one of myth elimination but rather one of theological demythologizing. This does not entail the elimination of myth; rather, the program involves interpreting myth for its views on human existence.15 This is crucial to Rudolf Bultmann, “New Testament and Mythology: The Problem of Demythologizing the New Testament Proclamation,” in New Testament and Mythology: And Other Basic Writings, ed. Schubert M. Ogden (Philadelphia: Fortress Press, 1984), 9. 11 Congdon, “Demystifying the Program of Demythologizing,” 8-10. 12 Bultmann, “New Testament and Mythology,” 3. 13 Bultmann, “New Testament and Mythology,” 4. 14 Rudolf Bultmann, “The Question of Wonder,” in Faith and Understanding (Philadelphia: Fortress Press, 1987), 247. 15 Bultmann argues that previous generations have gone astray in a partial elimination of myth from the New Testament. Bultmann argues this is wrong on at least two counts. First, myth should be examined for its interpretation of human existence and thereby not eliminated but demythologized. Second, the entirety of the New Testament presupposes a mythical world picture, even the depiction of the salvation occurrence. Therefore, one cannot “pick and choose” but must instead operate with a consistent program of demythologizing. See further Bultmann, “New Testament and Mythology,” 8-15. 10

Rudolf Bultmann (1884–1976)

89

grasp if we are to understand Bultmann’s theological engagement of the Bible’s claim that God is Creator. Cosmology must be interpreted in “anthropological terms—or, better, in existential terms”; it does not express an objective or scientific portrait of the world but instead our experiences “as the ground and limit of our world and of our own action and passion.”16 Affirming Faith in God as Creator in an Age of Science If the scriptural cosmologies—which are classic expressions of mythological thinking for Bultmann—do not give us a truthful account of the world, how is it that the modern Christian can affirm faith in God as Creator when the mythical world picture of Genesis 1–3 has become impossible? Modern science has, says Bultmann, “destroyed the old creation stories, even that of the Old Testament.”17 But this is no problem for the modern Christian since these creation myths are not intended to advance a scientific or even rational explanation for the origin of the cosmos and humanity’s place within it. Faith in God as Creator “is not a theory about some past occurrence such as might be depicted in mythological tales or cosmological speculation and natural scientific research; rather it is faith in man’s present determination by God.”18 And this faith in God the Creator is something that cannot be possessed as a piece of information or an item of knowledge; rather, faith in God as Creator must be appropriated and realized constantly in one’s life. Given that creation is “not an axiom of science under which the whole world process can be subsumed,” Bultmann says that he cannot “understand or ‘interpret’ something outside myself as creation of God or act of God. When I so speak, I am primarily saying something about myself.”19 Bultmann’s interpretation of the creation myths provides an insightful example of how his demythologizing hermeneutic works. For Bultmann, creation myths originate out of humanity’s anxiety in the world—namely, Bultmann, “New Testament and Mythology,” 5. Rudolf Bultmann, “The Meaning of the Christian Faith in Creation,” in Existence and Faith: Shorter Writings of Rudolf Bultmann (Cleveland, OH: Living Age Books, 1960), 209. 18 Bultmann, “Meaning of the Christian Faith in Creation,” 220. 19 Rudolf Bultmann, “The Question of Wonder,” in Faith and Understanding (Philadelphia: Fortress Press, 1987), 251. 16

17

90

Joshua W. Jipp

out of a sense that humanity is not the lord of its own existence and is subject to powers and forces out of its control. The creation myths originate, then, as an attempt to teach humanity something about its present situation—namely, that God continues to always be the source of humanity’s existence (e.g., Ps 104:30; Is 45:9-12; 64:8).20 Faith in God as Creator reminds humanity that it is not the lord of its world, that its very existence is contingent and dependent on powers greater than itself, and that the unpredictability of this world cannot be banished. Faith in God as Creator of this world reminds us that “human life is insecure; its course is not at man’s disposal. The man who is entrusted to himself does not have himself in hand. His life rests on the basis of a riddle, of the uncanny, and is constantly threatened.”21 Humanity further learns from the creation myths that God did not intend for it to be an isolated subject existing “from himself and for himself ” but has rather been created to be in relation to other humans.22 When humanity forgets that “he is himself from others and for others,” this demonstrates the reality of original sin that poisons our interpersonal relationships and leads to further insecurity and even hatred of our fellow humans.23 Bultmann finds similar theological reasoning in the apostle Paul and the Gospel of John. The apostle Paul, says Bultmann, invokes God as Creator, not as “a cosmological theory which professes to explain the origin of the world” but rather as “a proposition that concerns man’s existence.” In other words, the knowledge of God as Creator is in reality “knowledge of man . . . in his creatureliness and in his situation of being one to whom God has laid claim.”24 This dynamic is seen clearly in Bultmann’s explanation of Paul’s affirmation to the Corinthian church that for us, “even if there are so-called gods, whether in heaven or on earth (as indeed there are many “gods” and many “lords”), yet for us there is Bultmann, “Meaning of the Christian Faith in Creation,” 208. Bultmann, “Meaning of the Christian Faith in Creation,” 215. 22 Bultmann, “Meaning of the Christian Faith in Creation,” 216. 23 Bultmann, “Meaning of the Christian Faith in Creation,” 216-17. One should note here that Bultmann is significantly adapting traditional doctrines of original sin, which refers to the notion of human corruption (and, for some, guilt) that flows to all of Adam’s descendants due to the disobedience of Adam. 24 Rudolf Bultmann, Theology of the New Testament, 2 vols., trans. Kendrick Grobel (New York: Scribners and Sons, 1951–1955), 1:228. 20 21

Rudolf Bultmann (1884–1976)

91

but one God, the Father, from whom all things came and for whom we live; and there is but one Lord, Jesus Christ, through whom all things came and through whom we live” (1 Cor 8:5-6). Of course, we no longer believe in these many gods and lords, yet when interpreted rightly from the standpoint of myth, we see Paul is denying that the Christian can find his or her life in the so-called gods. In other words, wherever “the ultimate reality that gives meaning to our life and demands our worship is seen to lie in these powers, the many gods and lords still hold sway.” Bultmann says that these so-called gods and lords seek to exert power “in natural life and gives form to the nomos of nation and state, which, indeed, is frequently identified with it.”25 God is beyond every form of nationhood, history, art, and scientific discovery. God is their source. Thus, to affirm that God is Creator and that we are from him “means absolutely and in every present to have one’s source in him, in such a way that were he to withhold his creative will the creature would fall back into nothing.”26 Thus, the first article of faith in God as Creator is a faith that we, along with all of human history, are absolutely nothing. This is what the psalmist speaks of when he says, “When you hide your face, they are terrified; when you take away their breath, they die and return to the dust” (Ps 104:29).27 But this faith in God as Creator becomes Christian faith when we can further affirm that we exist from and for the Lord Jesus Christ (1 Cor 8:6). Bultmann here moves to the cross of Christ to explain Christ as Creator. To have faith in the crucified one means to permit oneself to be crucified with him, to permit this judgment also to be passed against oneself. To have faith in the cross of Christ means to be prepared to let God work as the Creator. God creates out of nothing, and whoever becomes nothing before him is made alive.28

For Bultmann, when we appropriate this understanding of God as Creator we are empowered to live in the world in such a way that we give tribute not to the so-called gods and lords but rather to God the Creator. Rudolf Bultmann, “Faith in God the Creator,” in Existence and Faith, 174. Bultmann, “Faith in God the Creator,” in Existence and Faith, 175. 27 Bultmann, “Faith in God the Creator,” in Existence and Faith, 177-78. 28 Bultmann, “Faith in God the Creator,” in Existence and Faith, 181. 25 26

92

Joshua W. Jipp

With respect to the Gospel of John, Bultmann rightly posits its major theme as divine revelation, and this is set forth emphatically in John’s prologue, which speaks of the preexistent Son as engaged with the Father in the act of creation (Jn 1:1-3). It should not be surprising, however, to hear Bultmann reject the view that John is concerned with cosmological or theological speculation about the origins of the world; rather, the prologue’s emphasis on revelation stems from humanity’s experience of new self-understanding in light of God’s revelation.29 John’s language of new birth (e.g., Jn 1:12-13; 3:1-16, 36) is ultimately a call for the individual to decide whether to “remain what he was—i.e., to remain in his old existence or not.”30 John’s remarkable cosmological dualism—with contrasts between light/darkness, truth/falsehood, life/death—is indebted to a pre-Christian Gnostic Redeemer myth (reconstructed by Bultmann out of the Mandaean literature). John has indeed taken this myth over, as evidenced through Jesus’ revelation discourses, but John himself has already begun the process of demythologizing as he engages in critical interpretation of the Gnostic myth and presses their dualism into an existential call for humanity to respond to revelation. In other words, the “cosmological dualism of Gnosticism has become in John a dualism of decision.”31 Primarily this is seen in the way in which the Fourth Gospel treats revelation historically and not cosmologically such that a material dualism gives way to an existential dualism of crisis that calls for humanity to a decision of faith.32 John’s Jesus engages in a proto-version of demythologizing, then, as Jesus reveals nothing about cosmology or human identity; he simply reveals that he is, in fact, the Revealer.33 God, Cosmology, and Science Bultmann’s legacy in the study of the New Testament as well as his hermeneutical contributions to the interpretation of the Scriptures have

Rudolf Bultmann, The Gospel of John: A Commentary, trans. G. R. Beasley-Murray (Oxford: Basil Blackwell, 1971), esp. 36-47. 30 Bultmann, Theology of the New Testament, 2:24. 31 Bultmann, Theology of the New Testament, 2:21. 32 See here especially, Bultmann, Theology of the New Testament, 2:26-40. 33 Bultmann, Theology of the New Testament, 2:66. 29

Rudolf Bultmann (1884–1976)

93

been studied and debated at great length,34 but only recently has Bultmann been described as a theologian of mission or even an intercultural theologian.35 In short, what is meant here is that Bultmann’s hermeneutical approach to myth emancipates theology from its captivity to any one particular culture, worldview, scientific theory, or cosmology. Bultmann has presented powerful and moving readings of the scriptural creation stories as they pertain to how humanity understands God as the source of its existence and how humanity has been created not for isolation but for interpersonal relationships. In one sense, Bultmann’s articulation of Scripture’s portrait of God as Creator nicely illumines how creation is invoked for anthropological purposes, and in this regard many of his insights are still timely and valuable. The biblical depictions of God as Creator are clearly not articulated by a dispassionate observer but are, indeed, concerned with God’s bearing on the world and human life. At its heart, however, Bultmann’s project enshrines within itself a set of strong dichotomies that many Christians will find unacceptable and unnecessary.36 While novelty is not in and of itself either a vice or a virtue, readers of Bultmann should recognize that his approach to science and myth is representative of an approach to the theology/ science dialogue that is far outside mainstream Protestant and Catholic theology. Markus Barth was not exaggerating, then, when he claimed that if Bultmann’s theological program was right, then “the church’s proclamation and New Testament studies over many hundreds of years must, totally and without further ado, do penance.”37 Two of Bultmann’s starting dichotomies strike me as innovative and problematic. First, Bultmann’s opposition to objectifying thinking within theology results See, for example, many of the essays in Bruce W. Longenecker and Mikeal C. Parsons, eds., Beyond Bultmann: Reckoning a New Testament Theology (Waco, TX: Baylor University Press, 2014). 35 I am referencing here the learned work of David W. Congdon, The Mission of Demythologizing. See especially chapter four. 36 I find especially helpful here the sage evaluation by Nils Dahl, “Rudolf Bultmann’s Theology of the New Testament,” in The Crucified Messiah and Other Essays (Minneapolis: Augsburg, 1974), 90-128, esp. 116-22; and Robert W. Yarbrough, The Salvation Historical Fallacy: Reassessing the History of New Testament Theology (Leiden: Deo, 2004). See also Gordon E. Michalson, “Bultmann’s Metaphysical Dualism,” Religion in Life 44 (1975): 454-61. 37 Markus Barth, “Die Methode von Bultmanns Theologie des Neuen Testaments,” Theologische Zeitschrift 11 (1955): 13. Quoted in Yarbrough, Salvation-Historical Fallacy, 283. 34

94

Joshua W. Jipp

in a science/faith dichotomy that I find unnecessary and unhelpful.38 Bultmann’s science/faith dichotomy would seem to result in science not contributing or illuminating anything about the Christian faith and, even more seriously in my view, has resulted in the view that God does not personally and actively interact with and on the world. God’s activity takes place only in the believer’s subjective and interior self-understanding. For Bultmann, to engage in ontological or metaphysical talk about God misunderstands the eschatological nature of revelation and inappropriately turns God into an object of human culture and knowledge. 39 Most Christians have affirmed that the Christian God is a personal being who creates, conserves, and rules the world.40 But has science actually destroyed these beliefs, thereby providing Bultmann with one cue for ­demythologizing the cosmologies in the Bible?41 Has science rendered obsolete the Bible’s constant claims that God acts within the world? Karl Jaspers argued that Bultmann’s hermeneutical program operated with a superficial understanding of science that exaggerates the surety and finality of its results as well as the differences between the ancient world and the modern.42 Similarly, Hans Jonas took Bultmann, his former teacher, to task for having “given more to modern science than is its

It seems to me that Kant’s critiques of reason, mediated and revised through Marburg neoKantian philosophers Paul Natorp and Hermann Cohen (who also influenced Bultmann’s teacher Wilhelm Hermann), are responsible here. See also Hans Jonas, “Is Faith Still Possible? Memories of Rudolf Bultmann and Reflections on the Philosophical Aspects of His Work,” Harvard Theological Review 75, no. 1 (1982): 12-15; and Roger A. Johnson, The Origins of Demythologizing: Philosophy and Historiography in the Theology of Rudolf Bultmann, Studies in the History of Religions (Leiden: Brill, 1974), esp. 38-84. 39 See here Congdon, Mission of Demythologizing, 372-74. Congdon views Bultmann’s project, however, much more positively: “Bultmann’s proclamation of an eschatological God is thus, finally, the proclamation of a missionary God. A God who is soteriologically transcendent and wholly other, whose being is in the ever new coming to the world, and whose saving act occurs in a forensic-eschatological kerygma is a God who is always in via to the new situations. Such a God cannot be objectified and so grasped as something available for observation and enjoyment” (374). 40 See Alvin Plantinga, Where the Conflict Really Lies: Science, Religion, and Naturalism (Oxford: Oxford University Press), 65-68. 41 Again, this is not the only reason for Bultmann’s demythologizing hermeneutic. But I do not think one can deny that it is indeed one very important reason for his program. 42 Karl Jaspers, “Myth and Religion,” in Myth and Christianity (New York: Noonday Press, 1958), 134-37. More generally, see Del Ratzsch, Science and Its Limits: The Natural Sciences in Christian Perspective (Downers Grove, IL: InterVarsity Press, 2000), 92-99. 38

Rudolf Bultmann (1884–1976)

95

due.”43 Modern science, says Jonas, does not engage in ontological claims. According to Jonas, Science merely says that for every occurrence one should seek a natural explanation until it is found, yet without endowing the laws of nature . . . with that kind of inviolability on principle . . . which only logical and mathematical rules enjoy. In other words, science issues a methodological command, not a metaphysical proposition.44

Bultmann repeatedly makes claims indicating his methodological commitment to a view of the world as governed by immutable natural laws, which must result in abandoning traditional notions of miracle and divine activity. Bultmann’s language regarding laws governing nature indicates he sees these laws as mechanistic, immutable, and necessary. But has modern science and the so-called laws of nature rendered impossible the traditional notions of God acting in the world to create and conserve? I am not a scientist, and I am sure there are a diversity of informed opinions, but I wonder whether Bultmann’s treatment of science seems fair for contemporary believers who are trained scientists, many of whom do not believe that science precludes wonder and miracle. New Testament scholar Luke Timothy Johnson argues that these so-called immutable laws of nature “are actually sets of hypotheses drawn from human observation” and are “far more fragile and uncertain than the unalterable laws that miracles are thought to disrupt.”45 In actual fact, assumptions regarding consensus in the natural sciences are “repeatedly disconfirmed by the history of science and sometimes by cultural history.”46 It may be wiser, in fact, as Alister McGrath has argued, to work to conceptualize the relationship between the natural sciences and theology as mutually illuminating disciplines—two disciplines and discourse that can enable us to “glimpse a richer vision of reality—not of course in the simplistic and long-discredited sense of forcing them into the same preconceived mould but rather in the more complicated one of Jonas, “Is Faith Still Possible?,” 9. Jonas, “Is Faith Still Possible?,” 10, italics mine. 45 Luke Timothy Johnson, Miracles, Interpretation (Louisville, KY: Westminster John Knox, 2018), 34. 46 Johnson, Miracles, 35. 43 44

96

Joshua W. Jipp

paying a respectful attentiveness to their distinct approaches.”47 For ­McGrath, in fact, it is precisely the truthfulness of Christian claims about God and the world that makes “sense of the successes and limits of the natural sciences.”48 Similarly, Alvin Plantinga gets precisely at this point when he notes that quite a few well-educated people (including even some theologians) understand science and history in a way that is entirely compatible both with the possibility and with the actuality of miracles. Many physicists and engineers understand “electrical light and the wireless” vastly better than Bultmann or his contemporary followers, but nonetheless hold precisely those New Testament beliefs.49

Further, Plantinga has presented a powerful argument that there is in fact a deep concord between science and theistic faith as it pertains to the image of God, the reliability and regularity of the world’s operations, human cognitive capacities, mathematics and law.50 In other words, not only is the supposed conflict between science and theism superficial but there are actually good, rational reasons for believing theism and science to be deeply compatible. And if science has not rendered impossible or obsolete the belief that God acts as a personal being within the world, then we may question whether the cosmological claims of the Bible are best interpreted solely within an existential and anthropological framework. Second, Bultmann is right that the scriptural creation texts are certainly deeply concerned with human life and existence. But does this mean that revelation is unable to engage in any meaningful or truthful talk of ontology, metaphysics, or cosmology?51 The biblical authors were almost certainly not interested in giving a scientific account of the origin of the world, but their concerns to say who God is, who and what Alister McGrath, Enriching Our Vision of Reality: Theology and the Natural Sciences in Dialogue (West Conshohocken, PA: Templeton Press, 2017), 183. 48 McGrath, Enriching Our Vision of Reality, 6. 49 Alvin Plantinga, “Two (or More) Kinds of Scripture Scholarship,” in Theology, History, and Biblical Interpretation: Modern Readings, ed. Darren Sarisky (London: Bloomsbury/T&T Clark, 2015), 395. 50 Plantinga, Where the Conflict Really Lies, 265-303. 51 One may give a positive answer here only if one has accepted Bultmann’s strict epistemological and ontological dualisms. Again, see Johnson, Origins of Demythologizing, 78-86. 47

Rudolf Bultmann (1884–1976)

97

humanity is, how humanity relates to animals and the rest of the created order, the institution of marriage, the nature of sin, and how God relates to the world seem to depend on our ability to engage in some articulation of the meaning of God’s past act of creation.52 In other words, knowing how we should live, who and what we are, the kind of world we inhabit, and the telos of human existence “depends in some measure on how we think the world is founded.”53 And this will, in my view, demand a deeper relationship between both science and theology, subject and object, and faith and worldview/cosmology. Bultmann believed that his demythologizing hermeneutic enabled one to encounter the truth of divine revelation, in part, by making it clear that one need not accept the mythical picture of the Scriptures as providing a primitive but still objectively truthful account of the world.54 The kerygma or gospel simply cannot be bound to stories, cosmologies, worldviews, creeds, or cultural artifacts. But this complete opposition to all forms of worldviews seems to me to come at too great a cost, one at odds with the New Testament’s advancing of the common confession that Jesus was Israel’s Messiah, crucified, buried, risen, and enthroned at God’s right hand. I have made it clear that I do not think Bultmann’s hermeneutical and theological approach to the doctrine of creation provides a viable way forward for those who want to affirm the traditional view of God as one who has acted and continues to act within the world, for those who want to affirm that the Bible’s cosmological claims do make meaningful ontological statements about the world we live in, and for those who see a more dynamic relationship between science and theology. But I would imagine that most who read Bultmann’s essays on God as Creator will find his analysis to be powerful in its articulation of the way in which these texts do indeed illuminate humanity’s deepest questions, hopes, and fears. Can we provide a better way forward, one that avoids Bultmann’s See, for example, Anthony C. Thiselton, The Hermeneutics of Doctrine (Grand Rapids, MI: Eerdmans, 2007), 198-207; see also Matthew Levering, Engaging the Doctrine of Creation: Cosmos, Creatures, and the Wise and Good Creator (Grand Rapids, MI: Baker Books, 2017). 53 Norman Wirzba, Food and Faith: A Theology of Eating (Oxford: Oxford University Press, 2011), 44. 54 On Bultmann’s opposition to all forms of objectifying thinking and the concept’s origin within Marburg neo-Kantian philosophy, see Congdon, Mission of Demythologizing, 369-74. 52

98

Joshua W. Jipp

strong dichotomies but continues to find in the Bible’s doctrine of creation a revelation that meets humanity’s deepest longings? Recommended Reading Primary Works Bultmann, Rudolf. “The Meaning of the Christian Faith in Creation.” In Existence and Faith: Shorter Writings of Rudolf Bultmann, 206-25. Cleveland, OH: Living Age Books, 1960. ———. “New Testament and Mythology: The Problem of Demythologizing the New Testament Proclamation.” In New Testament and Mythology: And Other Basic Writings, edited by Schubert M. Ogden, 1-43. Philadelphia: Fortress Press, 1984. ———. Theology of the New Testament. Translated by Kendrick Grobel. 2 vols. New York: Scribners and Sons, 1951–1955.

Secondary Works Congdon, David W. The Mission of Demythologizing: Rudolf Bultmann’s Dialectical Theology. Minneapolis: Fortress Press, 2015. Dahl, Nils. “Rudolf Bultmann’s Theology of the New Testament.” In The Crucified Messiah and Other Essays, 90-128. Minneapolis: Augsburg, 1974. Hamman, Konrad. Rudolf Bultmann: A Biography. Translated by Philip E. Devenish. Salem, OR: Polebridge Press, 2013.

5 KA R L BA RT H ( 1 8 8 6 – 19 6 8 ) The Doctrine of Creation and the World of Science K at h erine S ondere g g er Karl Barth descends from an extended family line of pastors and theologians, all in the tradition of the Swiss Reformation. While a young university student, Barth fell under the influence of the great architects of modern academic liberal theology, Immanuel Kant and Friedrich Schleiermacher. The young Karl Barth became a liberal in theology. Then something broke open. Barth began the long journey to his mature dogmatic theology. The first expression of his break with liberalism led to the “dialectical” period in his theology. The most celebrated work of this early phase was Barth’s Epistle to the Romans, issued in two editions, the first in 1919 and the more radical in 1922. Then in his early appointment in Göttingen, then later in Münster and Bonn, Barth made attempts at a new style of dogmatic theology, “church dogmatics.” After a false start or two, Barth began his lifework: Church Dogmatics. Published in four large volumes, it was a Christocentric theology, anchored in an “analogy of faith” that would occupy him throughout his long life and would be his task on the day of his death. At his death, The Church Dogmatics remained incomplete, with Barth’s Doctrine of Redemption, the proposed fifth volume, in only its barest sketch. Like the great cathedrals of Europe, Barth’s theology, though massive, was left in some high recesses unbuilt, yet even in its unfinished state it dominates the high ground and commands the theological skyline to this day.

T

he news from Karl Barth, it seems, is not good: the world of science appears to have no place at all in Barth’s massive Doctrine of Creation, the entire third volume of the magisterial Church Dogmatics. Here is Barth’s programmatic statement in the preface to his entire treatment of creation: It will perhaps be asked in criticism why I have not tackled the obvious scientific question posed in this context [of the doctrine of the work of

100

Katherine Sonderegger

the Creator.] It was my original belief that this would be necessary, but I later saw that there can be no scientific problems, objections, or aids in relation to what Holy Scripture and the Christian Church understand by the divine work of creation.1

Strong words! Now, this seems final enough, even ominously so, but Barth actually reserves some stronger comments for a bit further on in the preface: “The relevant task of dogmatics has been found exclusively in repeating the ‘saga’ [of Genesis 1 and 2], and I have found this task far finer and more rewarding than all the dilettante entanglements in which I might otherwise have found myself.”2 Then, with an air of repenting a bit of the evil he intended, Barth appends an alluring sketch of the two complex realms we explore in this volume, theology and natural science: There is free scope for natural science beyond what theology describes as the work of the Creator. And theology can and must move freely where science which really is science, and not secretly a pagan Gnosis or religion, has its appointed limit. I am of the opinion however, that future workers in the field of the Christian doctrine of creation will find many problems worth pondering in defining the point and manner of this twofold boundary.3

Now we might appoint ourselves some of these “future workers in the field of the Christian doctrine of creation” and so might find ourselves perhaps especially attuned to the methodological comments Barth offers here for the complex boundary between theology and science. They do indeed appear a bit more hospitable to our endeavors here, a sliver of light, perhaps, and might suggest themselves as the place to begin an expansion of Barth’s doctrine of creation in a scientific age. But I would like to begin our exploration of Barth’s account of this matter a bit further back, in his opening conviction—much firmer and more spirited, and more challenging—that the proper doctrine of creation belongs not to “dilettantish” comments about natural science but rather with extended reflection on the opening chapters of Genesis, the “saga of creation” as Karl Barth, Church Dogmatics, ed. G. W. Bromiley and T. F. Torrance, trans. J. W. Edwards, O. Bussey, and H. Knight (Edinburgh: T&T Clark, 1958), III/1:ix, emphasis mine (hereafter CD). 2 Barth, CD, III/1:ix-x. 3 Barth, CD, III/1:x. 1

Karl Barth (1886–1968)

101

Barth styles these sections. These comments are infinitely more concrete than Barth’s more general assessments of theology and science as intellectual fields—and concreteness is one of Barth’s most reliable laudatory terms—and I think they offer us a clear point of departure for a proper Christian reflection on the absolute origin of all things in God. But before I begin, the reader might benefit from a bit of background on Barth himself: Just who was he, and where does he stand in the long rank of distinguished Reformed theologians? Karl Barth belongs to that distinctive gift that the Protestant Reformation has given to the church: he descends from an extended family line of pastors and theologians, all in the tradition of the Swiss Reformation. Born in Basel in 1886, Barth retained, for all his cosmopolitanism, the particular thought world and habit of a Basler: he was at once bohemian and staunchly traditionalist. Young Karl was raised with a firm hand by a stern mother, Anna Katharina (born in the prominent Sartorius family), who influenced the grown Barth all his life, and by a father, Fritz, who taught New Testament and church history among the “positive theologians” at the University of Berne. While a young university student, Karl Barth fell under the influence of the great architects of modern academic liberal theology, Immanuel Kant and Friedrich Schleiermacher (more on him in a bit). Under his father’s watchful eye, the young graduate student traveled to the centers of advanced thought in German Protestantism: to Berlin and, above all, to Marburg. In the years before the First World War, Barth showed a marked interest in the historicism of Adolf von Harnack and the intense christological inwardness of Wilhelm Hermann, himself a Schleiermacherian of idealist coloring. The young Karl Barth became a liberal in theology. After his seminary studies he worked for Martin Rade’s celebrated journal Die Christliche Welt (The Christian World), a flagship of advanced liberal theology. Then, something broke open. Barth began the long journey to his mature, ecclesial, dogmatic theology. There are many ways this story of Barth’s break with liberalism is told—even by Barth himself! We will not go wrong, I think, if we assume that like all true conversions or revolutions, the causes are many and overlapping, and each cannot do the work of explanation on its own. Barth was a Swiss national in Germany during World War I; he was

102

Katherine Sonderegger

appalled at the liberal warmongering and the unvarnished nationalism of his one-time Christian socialist teachers. In those years, Barth’s own political commitments radicalized. He came under the influence of ­thoroughgoing religious socialists such as Leonhard Ragaz and Hermann Kutter, a pastor in Zurich. A young pastor in the industrial areas of Switzerland, the Aargau, Barth endorsed worker strikes and read widely in the more advanced radical literature of the postwar era. Barth was also a preacher and teacher: he knew that he must address his congregation in Safenwil, and his confirmands, out of the riches of Holy Scripture, and he believed his liberal training gave him not one step up in his attempt to do this. What has been called the “biblical realism” of the Blumhardts, father and son, left a strong mark on the young Barth: suddenly Scripture spoke to him as never before. But the conceptual and philosophical legacy of his academic training did not permit a direct, “pre-critical” reading of the Bible. Instead, Barth drew on his close study of the early Luther—Germany was in the grip of a Luther renaissance in those years— and of Soren Kierkegaard. This rich amalgam led to what scholars have called the “dialectical” period in Barth’s theology. What Schleiermacher held together in systematic relation—God, the world, and the self— Barth now pried apart. One could speak only of one element, then the other, but always in motion: dialectical, dynamic relation. The most ­celebrated work of this early phase was Barth’s Epistle to the Romans, issued in two editions, the first in 1919 and the more radical in 1922. Barth burst onto the scene, a theologian armed with a program and a groundswell of urgency. Liberalism would be overcome through dialectic and a stern renunciation of the analogy of being, a form of natural theology, as Barth saw it, dominant in both Catholic and Protestant liberal circles. Such a firestorm could not last. Barth, long before his allies and students, saw that a form of positive, or analogous, predication was demanded in a theology centered on the Word made flesh. First in his early appointment in Göttingen, then later in Münster and Bonn, Barth made attempts at a new style of dogmatic theology, “church dogmatics.” After a false start or two, Barth began his lifework: Church Dogmatics. Published in four large volumes, it was a Christocentric theology that would occupy him throughout his long life and would be his task on the

Karl Barth (1886–1968)

103

day of his death. Like the great cathedrals of Europe, it remains incomplete, yet in its unfinished state it dominates the high ground and commands the theological skyline to this day. Throughout his life, Barth remained the activist of his early pastorate in the Aargau. A leader in the Confessing Church, Barth wrote much of the Barmen Declaration single-handedly, and he remained an ally of Dietrich Bonhoeffer and other members of the German resistance, even when Barth was removed from his teaching post in Bonn for refusing to sign the Hitler oath. Barth’s relation to Judaism was complex, but he maintained a lifelong fidelity to the new state of Israel and to his early conviction that Christianity is built on the Law and Prophets, so that a Christian church that is hostile to Jews is betraying its Lord, himself a Jew, born under the law and obedient to it. Barth disturbed his postwar allies in his opposition to the Cold War and his willingness to affirm that the church can prosper even within communist states. Such a remarkable theologian will not be a simple man. Barth’s marriage was a sometimes difficult and most certainly a fraught one. Nelly Barth (née Hoffmann) was one of the young Barth’s confirmands; they married when he was twenty-seven and she a bare twenty. Nelly and Karl oversaw a large, exuberant household, with one daughter and four sons— and from the late 1920s forward, a secretary and close ally, Charlotte von Kirshbaum. Barth’s relation with von Kirschbaum was certainly an intimate one—Christiane Tietz is required reading for the anguish and moral struggle of this relationship—and it cast a shadow over Nelly Barth’s life and her children.4 Yet it would not capture the specificity or complexity of this relationship to characterize von Kirschbaum as “the second Mrs. Barth,” as is often heard in jocular quips. (It was certainly not a sexual relationship over these long decades, although it may have been in the early days. The possibility of divorce was explicitly aired in those early days but, significantly, never carried out.) Von Kirschbaum, at great cost to herself, allied herself with Barth, lived under his roof with Nelly, and did so until she was hospitalized with failing health and dementia. After Karl’s death, Nelly continued to visit Charlotte (Lollo, as 4

Christiane Tietz, “Karl Barth and Charlotte von Kirschbaum,” Theology Today 74, no. 2 (July 2017): 86-111.

104

Katherine Sonderegger

she was called in the family) every day in her nursing home. Barth was a man who needed companionship, friendship, and interlocuters all his life: he demanded a vis-à-vis. His intimacies were charged and difficult, yet he seemed to find them necessary to his very great work in theology. Barth is without parallel as a theologian of the church, most especially in the modern era. He swam “against the stream” time and again, and he nearly singlehandedly brought Protestant dogmatics back into the great tradition of the church, especially in its Augustinian heritage. But he was a human being too and, like us all, someone who stood in need of the grace of Jesus Christ, a grace he honored through his long and splendid Church Dogmatics. This is the man and thinker we meet when we turn to his exegesis of Genesis 1 and 2 and its grounding of the doctrine of creation. Barth in these early comments lays down two requirements for a proper Christian dogmatics: that the doctrine of creation concern itself principally with Holy Scripture and that this dogmatic exegesis be a “free science,” unrestricted but unaided too by scientific study of the cosmos. Barth adds richness to these points in his opening thesis statements to sections 40 and 41: “Creation,” he writes there, “comes first in the series of works of the triune God and is thus the beginning of all things distinct from God Himself. . . . The history of this covenant is as much the goal of creation as creation itself is the beginning of this history.”5 However objective and metaphysical all this sounds to our ears, Barth is quick and quite firm to insist that the doctrine of creation is principally a doctrine of faith: Barth claims that everything we see in the world about us, even its very external and stubborn material reality, is an object of trusting belief; we receive this truth even as we receive and hold dear our belief in the God and Father of our Lord Jesus Christ, the Creator of heaven and earth. The doctrine of creation, most basically, is an articulus fidei: as with the church, the sacraments, and the atoning death of Christ for us and for our sake—as with the being and mercy and triunity of almighty God himself—creation is the subject of Christian belief. Our world is no more an object of bare, neutral knowledge, Barth says, than Barth, CD, III/1:42.

5

Karl Barth (1886–1968)

105

is the kingly rule of Christ or his mighty resurrection from the dead. In truth, Barth radicalizes this common point. In the end, he argues, the proper and final answer to skepticism of all kinds, especially solipsism— that I alone am real in the universe—is to believe in God. The reality of the external world, everything from the singularity and the infinite expansion of the “Big Bang” to the most ordinary experience of a lecture hall and of a lecturer—all these are confirmed to us, made reliable and the bases of a humane life, by our belief in creation as the first way and work of almighty God. Barth draws together faith and history succinctly: “The purpose and therefore the meaning of creation is to make possible the history of God’s covenant with humankind which has its beginning, its center and its culmination in Jesus Christ.”6 The doctrine of creation, then, is thoroughly christological, and in just this way it is distinctly and inescapably historical. Thus, the schematic of the whole. But let me linger a moment, right now, on this little word, historical, as it will give shape to the contrast Barth draws with scientific accounts of the world. Even more, it chisels a sharp edge to what Barth means when he affirms creation as a work of God “in the beginning.” Barth adds a famous summary of this distinctive notion of history in the thesis statement I cited earlier: “Since it [creation] contains in itself the beginning of time, its historical reality eludes all historical observation and account, and can be expressed in the biblical creation narratives only in the form of pure saga.”7 Notice here that Barth strives to say something that appears to exceed any human speech: he speaks of a history that escapes historical observation, a chronicle of real events that cannot be measured by ordinary instruments of empirical reality, and an unfolding narrative of earthly events that are genuine, historical, and real just because they are pure saga, not history, or ordinary history, if we may use that phrase. Barth aims to develop a category here that expands our very notion of the real and the temporal. He believes there are events in our world— indeed, the central, revolutionary events in our world—that take place according to their own free and glorious calculus and in their own Barth, CD, III/1:42. Barth, CD, III/1:42.

6 7

106

Katherine Sonderegger

sovereignty fashion an environment, an objectivity and power and temporality, that is all their own: they seize reality and bend it (a favorite imagery of Barth’s) to their own ends. They do not break, but they do bend the laws of time, of materiality, and of history. This odd term, saga, aims to do all that heavy work. Barth will make extensive use of this term once again in his doctrine of the resurrection, a locus once more where divine revolution is inserted into our weary world and the very notion of reality itself is exploded and made new. We might say, in shorthand, that Barth’s use of saga extends and enriches his early insistence on the eschatological character of divine incursions in the world. Creation and re-creation, origin and consummation, divine freedom and divine presence: all these ways of God us-ward will escape any ordinary measure and will not take place according to our rules and habits and laws but will rather, in turn, ground them all. Barth’s doctrine of creation, that is, will follow the ancient pattern of the Hexaem­eron, a study of the six days of creation; it will characterize them as historical and as saga, and it will defend the whole as bearing down on and caught up in the sovereign mystery of Jesus Christ. These are Barth’s building blocks for an enormous, ambitious, and lengthy doctrine of creation—some two thousand pages over four massive volumes—and they will sustain a rich account of God’s ways and works with creatures apart from any scientific cosmology at all. Now let me insert some distinctions here before unfolding what I consider remarkable and lasting about Barth’s particular doctrine of creation. I want to draw a fine line between what I see Barth doing here, especially in his handling of time and of history, with what many Protestant academic theologians have done since the pioneering work of Friedrich Schleiermacher. In The Christian Faith, Schleiermacher sought to ground religion in a sphere all its own, a state of being more primal and, in its own realm, more sovereign than any of the human dimensions that compose our everyday lives. In famous words, Schleiermacher early on tells us that “piety is neither a doing nor a knowing but a feeling,” a lively sense of our own sheer dependence on Another.8 Note that for Friedrich Schleiermacher, The Christian Faith, ed. H. R. Mackintosh and J. S. Stewart (Edinburgh: T&T Clark, 1928), §3:5.

8

Karl Barth (1886–1968)

107

Schleiermacher, religion occupies a sphere, a dimension of human existence, that is marked off from all others. It stands as the still mediate point between a world convulsed by actions and reactions, by mastery and by searching, by undergoing and surviving and receiving what others will do to us. Schleiermacher was not himself chary of action or speaking or intellectual investigation—not a bit of it! He was a true cosmopolitan: a disciple of the arts, of music and poetry most especially; a translator of Platonic dialogues; an author himself of dialogues of high literary and conceptual distinction; a feminist and progressive thinker; a public intellectual, weighing in on the founding of universities and the distribution of their departments. He was an active churchman, preaching and lecturing constantly, it seems, and finding time in all this for deep friendships, deep loyalties to communities of all kinds. Schleiermacher did not imagine that religion was a form of reaction or of defensive protection against the encroachment of secular learning. Rather, Schleiermacher knew that religious piety would enter into every field and would take on many forms of doing and of knowing, of ethical disciplines and intellectual arguments in every area of life. But, Schleiermacher warned, religion was none of these things at heart. We are not in truth entering into the deepest core of religion when we compare doctrines or defend particular exegeses or dogmas or call on the faithful to bear their cross or to follow their Lord. No, Christian piety is instead an intense and rounded awareness, a whole illuminated globe of inward dependence on Christ, the source of all blessing and communion. Nothing can shake this. There is no teaching of the natural sciences, no discovery of the historians or archaeologists, no social or political analyses that can undermine, threaten, or even rival the Christian religion; it moves in its own sea. This supreme confidence gave Schleiermacher great freedom in his own historical work on the Scriptures; it allowed him to read the nineteenth-century naturalists without hesitation and showed him to be a citizen of the modern intellectual world, fully embracing its discoveries and highest aims. Consider, for example, our central focus here, the doctrine of creation. In the opening paragraphs of The Christian Faith, Schleiermacher offers a largely negative account of the teaching: “The religious consciousness

108

Katherine Sonderegger

which is here our basis contradicts every representation of the origin of the world which excludes anything whatever from origination by God, or which places God under those conditions and antitheses which have arisen in and through the world.”9 The aim, in this rather complex definition, is to set forth the doctrine of creation in such a way that the sole content of the doctrine can be found in what it “excludes”—that anything should exist outside the originary power of God. It is, Schleiermacher reports, a “negative” definition. The second half of section 40 lays out a theme that will govern much of the God-world relation throughout the work: God cannot be set over against or within the structures of the creaturely realm, on pain of making the Creator subject to the conditions of the creature. Such a stern restriction on the Creator-creature relation will make divine action and intervention in the world singular, strictly controlled, and exceedingly rare. God will relate to the world as Absolute Cause, but it is not clear that any other mighty works can be properly assigned to God from within the constraints of Schleiermacher’s conceptual schema. That The Christian Faith is primarily and principally a Christology makes Schleiermacher’s achievement all the more striking. He has placed at the heart of his dogmatic theology the incarnation of the eternal Son, yet he interprets the person of Christ in such a way that the code of proper dogmatic conduct, cited above, is not breached. (His delicate remarks on the distinction of generation and creation show his keen awareness of the difficulty he will face in Christology.) To understand such a Christology, then, the doctrine of creation is vital. In section 39, Schleiermacher warns us that true Protestant dogmatics will rule out “every alien element”—and by this he means the realms of natural science and of geophysical history. This is because theology is anchored in piety; it is, in the famous words of section 15, the deliverances of “religious affections set forth in speech.” If the doctrine of creation is to have any purchase on the devout Christian, Schleiermacher says, it must exhibit piety in its world-facing dimension. Creation must be a doctrine about dependence on God, nothing else. In a proper doctrine of creation, Schleiermacher tells us, we will understand everything Schleiermacher, Christian Faith, §40:149-50.

9

Karl Barth (1886–1968)

109

in the universe as dependent on God for its existence—not only objects (substances) but the very structure of reality, those elements Kant called the “forms of intuition,” space and time. For just that reason, the dependence Christians acknowledge is absolute. This is a rigorous and throughgoing conceptual demand. In section 40, for example, we find Schleiermacher cautioning us to avoid the temptation to imagine that creation has something to do with knowledge about the origins of our universe or the rise of creaturely species and kinds. These are not only “category mistakes”; they represent “alien elements”—a distortion and defilement of pure doctrine. Schleiermacher tells us that the Old Testament contains the “beginnings of a history-book which as such chiefly satisfies the desire for knowledge”—not a commendation! All this may sound rather cheap, but Schleiermacher has some more deft and delicate analyses to offer too. He notes that “Hebrews” have their “better days,” in which creation is spoken of in broad brush strokes, and nothing much is insisted on beyond the confession that all things come from God’s provident hand. The book of Job, the Psalms, and the Platonizing exegete Philo all come in for praise on this score. Schleiermacher sees that the “Mosaic narrative” is not history in the ordinary sense; Genesis 1 does not record events in the way that 1 and 2 Samuel do. Yet Schleiermacher does not have enough dogmatic interest to investigate just what sort of history it might be. That task will be left to Barth. But he was a sensitive interpreter of ancient texts, and he knew that something distinctive could be found in these opening chapters of the Holy Book. Still, for all that, the point to be made is the plain one stated above: a matter of knowledge or historical detail cannot be an article of faith. Only the “higher consciousness,” piety, can be that. In such a theology, many traditional elements of the doctrine of creation must fall to the wayside. Luther and Calvin believed that Genesis 1–3 were historical accounts, but they were wrong. Like most modernist interpreters, Schleiermacher saw a distinct alteration between the style and aim of Genesis 1–2, the so-called two creation narratives of Genesis. From such distinctions grew the famous “documentary hypothesis” of nineteenth-century German higher criticism. For Schleiermacher, this marked difference between the chapters means that “we can hardly

110

Katherine Sonderegger

attribute to them a genuine historical character.”10 We glimpse here something of what German belle-lettrists considered history to be: a continuous, coherent, and self-consistent narrative of events. But Schleiermacher is quick to add that even should we be persuaded that Genesis 1–2 contains genuine or scientific history, we must sternly exclude it from our doctrine; it cannot be the basis of a belief in creation. In the same way, the traditional cosmological teaching of creation from nothing— creatio ex nihilo—cannot be considered a proper element of the doctrine of creation. True, everything must be originated from God’s absolute ­causality, but this is a deliverance of pious feeling, not a metaphysical principle. So too we must rule out the traditional conviction that the universe has a beginning in time. It is not simply that time itself is a creature— Schleiermacher affirms this—but the medieval debate, stemming from the works of Aristotle, over whether creation is eternal can find no place in a well-ordered doctrine of creation. Moreover, a doctrine of divine decrees cannot enter into the doctrine of creation, for this introduces a note of deliberation, an element of temporality, within the Godhead; and just this, Schleiermacher holds, makes God a creature, subject to the forces and oppositions of creaturely life. (Such reflections will draw Schleiermacher very close indeed to Spinoza, and this charge of Spinozism in the doctrine of God will dog Schleiermacher all his life.) Were all this not enough, Schleiermacher will dare a step further. He argues that in our inwardness we are aware simply of our finitude, our utter need of God. But we are not aware of our beginning. Rather, we experience only our continuance, our life unfolding before us. By Schleiermacher’s account, then, the doctrine of creation is more truly a doctrine of preservation: everything rests on God. The aim of this elaborate restriction in the doctrine of creation is intellectual freedom: the piety of the believer must not interfere with scientific or historical research. Schleiermacher writes, “The development of our self-consciousness must not be so conceived as to set the man who desires knowledge in contradiction with the principles of research he follows in the sphere of nature or of history.”11 By axiom, the religious Schleiermacher, Christian Faith, §40.2:151. Schleiermacher, Christian Faith, §39.2:149.

10 11

Karl Barth (1886–1968)

111

feeling of a Christian cannot conflict with rational, scientific inquiry, and a restriction on free investigation by a Christian would be a violation of the deepest principle of piety. Schleiermacher imagines a great calm breaking out over the anxious conflict between Scripture and historical criticism, between theology and natural science. A kind of dogmatic surgery will cut away the remnants of natural science embedded in doctrine, so that nothing of the old conflict can possibly reinfect the teachings of the church. Schleiermacher solemnly relinquishes possession of the natural world to science: The complete separation of these two [science and theology] involves our handing over this subject to natural science, which, carrying its researches backward into time, may lead us back to the forces and masses that formed the world, or even further still. On this assumption we may patiently await the result.12

By the 1830s, when the second edition of the Glaubenlehre made its appearance, naturalists were uncovering an extensive fossil record, and the geological age of the earth was being extended well past the four thousand years Bishop James Ussher assigned to it. So impressive was this work to an intellectual of Schleiermacher’s scope that he imagined scientific research uncovering the principles and constituents of the earth and, even more, of the universe itself. Modern physics would not have disappointed him. In every disputed question of his age—in the proper disciplines of the new universities; in the place of women within civic society; in the rise of paleontology and geology; in the histories of ancient societies, including those of the Bible—Schleiermacher showed a wonderful confidence and Christian élan. Nothing could touch piety; everything could be welcomed. Of course, we must take the measure, too, of the ground ceded in such a pacific victory. As with many freedoms, this one came at a cost. The essence of the Christian religion, for Schleiermacher and for his many followers, could not be found where the essence of the natural sciences, cosmology, human philosophy, or history lay. They occupied different worlds. There could be no competition, because the players stood on Schleiermacher, Christian Faith, §40.1:150.

12

112

Katherine Sonderegger

separate fields—even on distinct and autonomous planes—and they did not speak the same original tongue. The Christian religion did not lay claim to the external world as a whole, nor to human civilization as a history and a task; these were forms of knowing and doing that by rights belonged to other human enterprises. Piety was the Christian homeland. Now Schleiermacher has many disciples in our modern era. Those who consider science a realm apart (we might think of some of Stephen Jay Gould’s work here), a separate “magisterium” or distinct terrain, as does Daniel Dennett’s with his, perhaps ironic, “scientific,” “spiritual,” and “philosophical” levels are all disciples of Schleiermacher’s bold design. We often hear these days of “readings” that occupy differing linguistic and conceptual rooms (science in one, theology in another) or differing objects of investigation. These are nonaggression pacts, if I may style them so, and they owe their conceptual armature, if not their larger spirit, to the pioneering work of Schleiermacher’s Christian Faith. But we may well ask, Can Christian theology remain a field that encompasses and concerns the real if it demands no assent from us about the state of the natural world or the historical events that unfold on it? Inwardness, certainly, is to be prized in Christian thought, and a theology that does not invoke and rest on warm piety can hardly win our trust. Still, cosmology, metaphysics, natural history, natural law: all these have been instruments in the hands of our ancestors in the faith, elements of theological reflection. They expressed a confidence that Christian theology had to do with the universe we encounter, with all of it, and that doctrine could lay claim to world-encompassing realism. The ability to be wrong, Karl Popper famously said, was a mark of the true. Something that can never be proved wrong can never be right. We may not wholly agree with Popper here—but his instinct about the credibility and intellectual strength of a conceptual system is one Christians must lay to heart. To bequeath the natural world to natural science, as did Schleiermacher so effortlessly, is to reserve for the heirs of the faith a small plot of sacred land, an island of piety, that is peaceful, to be sure, but hardly noticed by the bustle and achievement of the larger intellectual world. Our contemporary intellectual society, with

Karl Barth (1886–1968)

113

its pronounced secularism, is the descendent of this peace accord—in its gains, certainly, but also in its losses. We have reason, then, to wonder whether the great alternative to The Christian Faith, Karl Barth’s Church Dogmatics, has left us a similar legacy. It may seem that Barth, too, follows Schleiermacher’s lead, at least in setting apart theology from scientific study. But here I think we want to be cautious: not everything that sounds the same is the same. Barth has clearly signaled here that he is dispensing with a scientific cosmology and with direct conversations or quarrels with astronomy, quantum mechanics, and evolutionary biology; he thought at one time he may have needed to do so but now knows nothing of the kind is needed or wanted. But here is where careful distinctions will come to our aid. Barth has something quite different in mind here than the separate realms of Schleiermacher’s peace accord with science. Barth instead is laying claim to the whole of reality, the whole metaphysical and historical cosmos, and placing it under the sovereignty of the Creator God of Genesis. Barth argues at length and with great technical sophistication that the world we and all scientists inhabit and study is a world of history, of pure saga and not of myth: that is the distinctive path Barth marks out as his own here. Barth shows little openness to segmenting reality— his willingness to speak of a two-sided border between theology and science does not, I think, violate this rule—and does not in the long exegesis in CD III/1 seek a separate realm for this reading nor a separate object to which that reading refers. Unlike much modernism in the field of religion and science, Barth does not hold that Christian theology refers to special domains within the world of creatures and of time: he is not one, for example, to speak of “salvation history” as if that would refer to events marked off from the history that all other peoples and nations experience and is known only to faith. No, Barth’s concepts of saga and of prehistory belong firmly to the one world we all inherit, and they characterize and explain the entire cosmos as creature, the free action of the free God. What the doctrine of creation undertakes in Barth’s view is a recasting of the whole of reality under a relation, the God-world relation. And that, in Barth’s judgment, expresses and enacts a form of history.

114

Katherine Sonderegger

We have now arrived at the fundamental conviction that governs Barth’s entire doctrine of creation—really, in truth, his whole Church Dogmatics. So deep and lasting is this conviction that we can find it lodged in Barth’s earliest essays and commentaries—consider his remarks on history and historiography in his second Letter to the Romans— and we can trace its presence throughout the long course of the Church Dogmatics, announcing itself with noisy triumph in Barth’s doctrine of reconciliation: the history of Christ’s incarnation, passion, and victorious rising. The relation between almighty God and his creatures belongs to a particular kind: it is not inert or timeless, not static or mechanical, but is rather the kind of relation we would call alive, a living, dynamic, and unfolding relation between two subjects, standing face-to-face, encountering one another across the infinite divide that is Creator and creature. Such a relation is an event; it takes place over time and in just this way has a narrative, a beginning and an ending. Simply said, the relation between God and creature is a history. Now, this is a peculiar sort of history—Barth is very straightforward about that. One way to think about this unique form of narrative is to consider a human-scale analogy, the analogy of friendship. This is obviously a relationship, one we call “personal”: a friend is someone we share our personal and private life with. It is not formal, as is a business partnership, say; it does not follow prescribed rules and patterns, as does a legally binding relation such as employment; and it lays claim to our whole selves—if it is a deep and lasting friendship—as do few other ties in our lives. What grows up between two friends is what Barth would call “history.” It is not simply what takes place (a dinner shared, a trip together, an illness overcome) but is rather the elaborate exchange that is a life shared—the physical, emotional, social, and spiritual dynamism that knits lives together. Barth holds that creation is history in just this sense, and in that way Barth takes his place among the prominent personalists of the interwar years. All of this must seem rather commonsensical and perhaps rather tame. But notice what is missing here. There is no metaphysical “fact of the matter” about a personal relation of this kind. We cannot submit friendship to a medical or mechanical test to measure its presence or

Karl Barth (1886–1968)

115

power. There is no object that is the friendship; we do not look for a particular thing and say, “See, here is the friendship.” We reserve the term symbol for an object that bears or represents, that stands in for the complex reality that is a living relation between persons. A ring does this for many marriages; a gift often performs this service for a friendship. But the object is not the relation; nothing is that. Nor does friendship follow a definition of the relation that governs and grounds all else. Friendship is not simply one thing, one relation, one pattern or event; rather, it is legion. A rich friendship follows its own arc—at times quiet companionship; at others an adventure or something a bit foolish; at times emotionally demanding or intense, a regret or anger, a sorrow, perhaps; and at times a reliable source of rides to the bus station or back home again. The salient metaphysical point here is that a living relationship does not exist as a tertium quid, a third reality next to the partners in the relationship; rather, it simply is the partners in this particular and concrete relation. Notice that this does not make the relationship irreal—not a bit of it! Not everything real is material or objective, nor is the real only measurable or substantial. Rather, the real may be a “state of affairs” (marriage is real in just this sense) but may also be simply two things in living relation, the happening between them that is their relation. Just so, Barth says, the very idea of creation, and the covenant that will spring out of its midst, is not a “fact of the matter” to be investigated or measured or tried, nor is it a feature or property of the cosmos to be picked out and examined per se. No, creation, like covenant, is simply the world, this world and all that is in it, in personal relation, living dynamism, with the Creator and Lord of the covenant. It is a form of friendship in just this way, a history of encounter. Now notice how strange such an encounter must be. We can imagine, mutatis mutandis, a friendship that just is the history between two people, but we cannot in truth imagine a relation or an encounter that is between the Lord God Almighty and a creature or every creature. One of the partners in this friendship escapes all our definitions and familiar understanding: almighty God is Lord. It is of course true that God is himself invisible and immaterial; he is not seen among his creatures as they are, and his Spirit is a relatio, a communion and friendship that just

116

Katherine Sonderegger

is God, and expresses the love who is Father and Son in the Godhead. Nor should we imagine that the relation will be plainer or more comforting in the incarnate Word, who is God visible among us. For we creatures experienced him as Lord, as master of sea and wind, as commander of demons and tormenters or every kind, as a teacher, yes, but One who filled his disciples with fear and silenced his opponents so that none dared to ask him anything more. This Immanuel offended us and offends us still. He is inconvenient to us, and his ways are strange. The cross is the name of the God-world relation under the conditions of sin: we are not friends of such a Lord but rather his persecutors. The very odd relation between God and creatures, that is, does not spring only from the majesty and invisibility of almighty God, though to be sure it concerns these. But it springs also, and most tragically, from the history that is ours, the primal rebellion that has its start in Paradise and follows us in every step from Eden to Golgotha, and from there to the violence against God in every age. The history of encounter between God and creation is also a history of sin. The narratives of Genesis 1–2, then, do not stand alone; they also constitute Genesis 3. History, Barth tells us, begins in Paradise, and in this very special sense Barth affirms the historicity of the events and figures of the Garden of Eden. But note this very special sense. This primal history unfolds the distinct and concrete relation between God and his creatures, and as such it will be a history unmeasurable by scientific historiography. It will take place. It is an event, and it belongs to the realm of time, creaturely time. Barth is very firm about this. Departing from centuries of commentary tradition, Barth does not consider “time” a dimension created along with the other elements of the six days. Rather, he boldly affirms that creation is the first work of God, and it is done in time, as the head of all other works of the Creator. This point, although surely a subtle, perhaps overly subtle, one on Barth’s point, matters to Barth’s program here, because the world is never timeless, never an element in an eternal relation that is pure origin, say, or pure causal power. Rather, creation from its very start is temporal, the history of the encounter of the Lord with his servants. This history, then, will be told in a unique form and in a unique voice. It will be “pure saga.” For Barth, this

Karl Barth (1886–1968)

117

term connotes a “divinatory and poetic imagination” that speaks of what cannot properly or directly be spoken of; it tells what has taken place in the only idiom suitable for the task, an extraordinary diction of vivid detail, mighty works no one could witness, repetition and solemn cadence, narrative mastery and irony, all in a world readily recognizable as our own. No legendary beasts, no symbolic talismans, no dismemberment and birth of the gods, no just-so stories such as the Symposium’s androgyne. Rather, Genesis is saga as the history of our world, filled with the creatures we know, behaving under these remarkable conditions yet in a way we consider “realistic” and “timely.” We do not look for some esoteric key that will unlock these narratives, nor do we expect that some conceptual secret lies behind these events, a riddle that is exploded and explained by adepts who have studied these mysteries. No, Genesis recounts the creation as a story about God’s creating the heavens and the earth and beginning the ways of the covenant with the human creatures he has made. Its meaning is its telling, and that makes it history, Barth says—the unfolding of events in time and in our world. Creation, then, cannot be a particular or formal relation; it cannot be, Barth says, a “timeless” connection or exchange. Creation, for example, cannot be simply causal: it cannot consist in an impersonal and atemporal relation such as that found in a Prime Mover and the moved, or in a Prime Cause, itself uncaused, and the worldly effects. Creation cannot be simply “dependence,” as if the hallmark of the created order were a property, good now and always, that marked out the physical as the creature of the Creator God. Nor could creation consist in a kind of spiritual effusion, an inner light that is a spark from the great Originary Fire: creation is not an inner metaphysical trait or essence that is handed over by or emanated from a Vital Source. No, creation is not a thing, not an object or property that belongs to the world. Most certainly, it is not a relation that timelessly establishes such a property. That is because in Barth’s view, creation is saga, not myth. A myth, Barth says, parodies a saga: it appears to need telling, recounting, and appears to involve various living beings in relation to one another, and it seems to concern the origin of the world we all inhabit. But in truth, Barth says, a myth only parodies these traits; in

118

Katherine Sonderegger

fact, it is none of these things. A myth takes many forms, and Barth often has in mind the cosmogonies of the ancient worlds of Babylon and Egypt, but also of Greece and Rome and of the Hellenizing Gnostics; but a myth can be altogether modern and scientific as well. The point in common here is that a myth relates a timeless relation: it does not take place, Celsus said, but always is. According to Barth, once we see this, the narrative elements of the mythos drop away, and we learn to see through them to these timeless truths, the primal cause, primal fertility, or primal struggle that is cosmic light and darkness in the world from beginning to end. Nothing truly happens here, Barth says; we enter instead into a realm of eternal return, a timeless cycle where everything returns to its beginning and the structure of reality is laid bare. Here we might ask our final question to put to Karl Barth’s doctrine of creation: Might science itself be a kind of mythos? Might it present a timeless, mythic relation of God to the world? Here, it seems to me, is the final contribution that Barth’s theology might make to the work we have before us here in this volume. What kind of science is proper partner to dogmatics? What kind of science is instead a “pagan Gnosis,” as Barth put this point in his opening remarks in the preface to volume three of the Church Dogmatics? Could we see the complex relation between theology and science ordered by a form that is itself a kind of history, a kind of saga that is not timeless but very much temporal, concrete, and anchored in this world of covenant and of reconciliation? What would that look like? I think we can all imagine in one way a science that has become mythic or pagan in Barth’s eyes: that is the science we have learned to call “scientism,” the timeless conviction that there is no God to create a world and that science alone measures the reality that is left. I think Barth would consider such covert forms of secular philosophy a kind of mythos in which the conceptual truth to be learned in the world is that we are here alone and that this infinitely expanding universe will one day dissipate and nihilate itself, even as it began, spontaneously and without purpose. This Barth calls a “world view,” and such a scientistic worldview cannot find a friendship with theology, because it is a secret dogmatism all its own. But might there

Karl Barth (1886–1968)

119

be another form of astronomy, cosmology, and biology that would exist fruitfully along the border of this world of saga and time? I believe that there might be—though I think this would have to be speculative on my part. Such a scientific account of the world would not simply tolerate theology or exist with it side by side without jostling or rivalry; rather, it would be companion to a world that has encountered its Lord. Its concerns would be for all physical realities, quanta as well as the very great, complex mysteries that inhabit deep space. That is, such a science would find room for medium-sized objects, for groups and herds, for flocks and families, if only as the emergent or upper-level objects that compose a Newtonian world. And it would lay out its theories with a place for temporality, not simply the B series and its tenseless relation of before and after, but a genuine openness to the A series, the world built up from past, present, and future. Such natural sciences would study the world we all inhabit; it would be neither reductive nor hostile to the world of purpose and of covenant that makes human history humane. It too would honor the inwardness of the human species and of its scientists, such that the natural sciences would make sense of the world that conscious beings such as ourselves order and undergo. It would be no more amoral than it would be atemporal, but it would instead be the intellectual work and discovery of life forms who also stand before their Lord and Maker, moral and historical beings all. These sciences would not study or produce saga or even history in the ordinary sense, but they would be fields that could underwrite or prepare for a historical reality of the kind dogmatics know. Even as creation is the external ground of covenant, so such sciences could be the deep structure and dynamism of the material and cultural world we see before us. Perhaps such a science could be, and perhaps—indeed, I am bold enough to hope this— it is the only science that has ever been. There is one world and one truth, and the Lord God of all creation is the origin, the ratio, and the scientia of all that is. Such is the science I believe Barth would have welcomed, and such is the science that is the glory of the real, and in just this way, the glory of the most real God who is the one Creator of heaven and of earth.

120

Katherine Sonderegger

Recommended Reading Primary Works Barth, Karl. The Doctrine of Creation. Vol. III/1 of Church Dogmatics. Edited by G. W. Bromiley and T. F. Torrance. Translated by J. W. Edwards, O. Bussey, and H. Knight. Edinburgh: T&T Clark, 1958. ———. The Epistle to the Romans. Translated by Edwyn C. Hoskyns. London: Oxford University Press, 1933. Schleiermacher, Friedrich. The Christian Faith. Edited by H. R. Mackintosh and J. S. Stewart. Edinburgh: T&T Clark, 1928. ———. Christian Faith: A New Translation and Critical Edition. Edited by Catherine L. Kelsey and Terrence N. Tice. Translated by Terrence N. Tice, Catherine L. Kelsey, and Edwina Lawler. 2 vols. Louisville, KY: Westminster John Knox Press, 2016.

Secondary Works Busch, Eberhard. Karl Barth: His Life from Letters and Autobiographical Texts. Translated by J. Bowden. Philadelphia: Fortress Press, 1976. Hunsinger, George. How to Read Karl Barth. New York: Oxford University Press, 1991. McCormack, Bruce. Orthodox and Modern: Studies in the Theology of Karl Barth. Grand Rapids, MI: Baker Books, 2008. McTaggert, J. M. E. “The Unreality of Time.” Mind 17 (1908): 457-74. Robinson, J. M. ed. The Beginnings of Dialectical Theology. Richmond: John Knox Press, 1968. Tietz, Christiane. “Karl Barth and Charlotte von Kirschbaum.” Theology Today 74, no. 2 (July 1, 2017): 86-111.

6 T. F. T O R R A N C E ( 1 9 1 3 – 2 0 0 7 ) Christ the Key to Creation and Theological Science Kevin J. Vanh o oz er Thomas Forsyth Torrance was born in China to missionary parents. He studied for a time with Karl Barth and later oversaw the translation and publication of Barth’s Church Dogmatics into English. Torrance served for twenty-seven years as professor of Christian dogmatics at New College, the divinity school of the University of Edinburgh, Scotland. He was a prolific author, writing books on the nature of theology, the history of hermeneutics, the relation of theology and science, and specific Christian doctrines like the Trinity, incarnation, and atonement. He was also founding editor in 1948 of the Scottish Journal of Theology. Torrance is widely recognized as the greatest British Protestant theologian of his generation and one of the most important Reformed theologians since Karl Barth. In 1978, Torrance received the Templeton Foundation Prize for Progress in Religion for contributions to the emerging dialogue between science and theology. He was an active churchman throughout his life and served as the moderator of the General Assembly of the Church of Scotland in 1976. In addition, he was a representative of Reformed churches in ecumenical dialogues with other churches, including talks with the Eastern Orthodox church on the doctrine of the Trinity.

C

hristian theologians should never back down when the topic is reality. If there is no God, and if Christ had not been raised in space and time, then as Kierkegaard says, “good night to Christianity.”1 T. F. Torrance’s commitment and openness to reality—the reality of God revealed in Christ and the reality of the physical universe created through Soren Kierkegaard, “The Difference Between a Genius and an Apostle,” in The Book on Adler, ed. Howard V. Hong and Edna H. Hong (Princeton, NJ: Princeton University Press, 1998), 173.

1

122

Kevin J. Vanhoozer

and for Christ and held together in him (Col 1:15-16)—was arguably the lodestar of his career, inciting him to go the second mile (often at the speed of light) to bring theology and the physical sciences into mutually edifying conversation. In defending theological realism, appropriating Einstein, and refusing to cede the rubric “science” to the natural sciences, Tom Torrance was both prophet to modernity and apostle to the physicists. He was also a remarkable synthetic thinker who perceived profound connections between patristic theologians and modern physicists. While, at first glance, ancient Alexandria might seem light years removed from Albert Einstein’s theory of relativity, when we understand why it is not, we begin to understand Torrance’s distinct insight and unique contribution to the theology/ science dialogue. Torrance represents a minority report in the contemporary discussion, in which theology figures no longer as the queen of the sciences but only as their poor sister—or worse, no relation at all.2 The single most important thing to grasp about Torrance is his understanding of science (and this includes theology as well as natural sciences like physics and biology) as the humble, attentive, yet disciplined way in which we come to know a particular object. It is not a matter of following a uniform step-by-step procedure, as if human methodology could harness reality. On the contrary, the particular subject matter in question dictates the method(s) for coming to know it: “The means to understanding must be in accordance with the substance of what is sought.”3 In brief, ontology (what there is) determines epistemology (how it is known). This is Torrance’s “fundamental axiom,”4 and it is a conviction that John Philoponus of ancient Alexandria, Karl Barth, and Albert Einstein all share. Let me introduce a shorthand term for this Torrance considers Clement of Alexandria “one of the first to attempt to put Christian theology upon a scientific basis.” T. F. Torrance, “The Implications of Oikonomia for Knowledge and Speech of God in Early Christian Theology,” in Oikonomia: Heilsgeschichte als Thema der Theologie: Oscar Cullman zum 65 Geburstag gewidmet, ed. Felix Christ (Hamburg: Herbert Reich, 1967), 223. See also Alister McGrath: “Torrance argues that Alexandria became the center of an approach to scientific method which was of considerable importance both to the development of the natural sciences and to Christian theology.” Alister McGrath, T. F. Torrance: An Intellectual Biography (London: T&T Clark, 1999), 210. 3 Daniel W. Hardy, “T. F. Torrance,” in The Modern Theologians, 3rd ed. (Oxford: Blackwell, 2005), 167, emphasis original. 4 Elmer M. Colyer, How to Read T. F. Torrance: Understanding His Trinitarian & Scientific Theology (Downers Grove, IL: InterVarsity Press, 2001), 322. 2

T. F. Torrance (1913–2007)

123

axiom: kataphysics. Unlike speculative versions of metaphysics that proceed from reason alone, a priori (apart from experience), positing certain first principles of reality, every true science is for Torrance kata physin—literally, “according to the nature (of the object)” and hence a posteriori (on the basis of experience of the object). Torrance sees a deep kataphysical connection between theological science (the knowledge of God) and the natural sciences (the knowledge of the physical world). The subject matter of theology is the being and work of the triune God, and theology is scientific insofar as it lets its subject matter disclose itself in its own way, through God’s self-disclosure in Christ through the Holy Spirit. Theology and the natural sciences stand in a positive, not antagonistic, relation: each adheres in its own way to the scientific method (kataphysis), and each is interested in the material world, although they yield knowledge of different aspects of the world. To anticipate, physics can measure the speed of light, but only theology knows light as created in and through Christ. How did the person responsible for translating Karl Barth’s Church Dogmatics into English come to be interested in the natural sciences, particularly given Barth’s own antipathy to natural theology? There is an answer, but to appreciate it involves baptism by immersion, not sprinkling, a fitting move given the way Torrance thinks we come to know things in general—less by ratiocination than by personal participation. In that spirit, I offer this essay in aid of our indwelling of (but not drowning in) the deep end of Torrance’s theological pool. It is an intimidating prospect: George Hunsinger’s How To Read Karl Barth comes to less than three hundred pages, but Elmer Colyer’s How to Read T. F. Torrance weighs in at four hundred. Deep waters indeed.5 A Kataphysical Poet: Knowledge of Reality in Torrance’s Theology Torrance is a big-picture thinker. Like Einstein, he is after a kind of unified field theory, a metanarrative that ranges over developments in I also found sobering a review of a book titled Incarnation and Physics: Natural Science in the Theology of Thomas F. Torrance (Oxford: Oxford University Press, 2002), which began by chastising the author for thinking that he could grasp Torrance’s thought after only two years’ research.

5

124

Kevin J. Vanhoozer

science, philosophy, and theology alike. Paul Molnar quips that Torrance “can summarize a thousand years of thought in one paragraph.”6 The present chapter presents Torrance as a “kataphysical poet”7 who seeks to penetrate into the nature of things—their “onto-relations”—via insight and intuition rather than observation and ratiocination.8 To be a kataphysical poet has nothing to do with producing creative fiction or (contra Kant) with imposing subjective modes of thought onto objects of perception. On the contrary, it is about being open and receptive to the disclosure of things outside oneself. This, at least, is what Torrance says he learned from his study of patristic theology, and Athanasius in particular (see below): “The material content of knowledge must be allowed to determine its form and method . . . knowledge in any field is governed by the nature of its object as it is progressively disclosed to us.”9 Torrance finds other theologians who share Athanasius’s kataphatic insight, and the story Torrance tells about the history of theology is basically a set of variations on a kataphysical theme. Like all stories, it has both its villains and its heroes, but it might be helpful to begin with some background regarding the conflict that lies at its heart.10 On poetics: Forms of speaking, thinking, experiencing, and being. To fully appreciate what Torrance has to say about theology and science requires poetics: the study of the ways in which forms of being, speaking, thinking, and experiencing relate to one another (or fail to relate). Torrance himself does not use the term poetics, yet he unfailingly calls our attention to the importance of how the forms that inhere in reality relate Cited in Douglas Farrow, “T. F. Torrance and the Latin Heresy,” First Things 238 (December 2013): 26. 7 This in comparison to and contrast with the so-called seventeenth-century metaphysical poets like John Donne. The label metaphysical poet, coined by John Dryden, was meant pejoratively to describe their stylistic and intellectual complexity. In contrast, kataphysical poet has no negative connotation. 8 Everything that exists is related to its Creator, who is personal. To think about anything real is thus to think about its onto-relations (i.e., its being in relation to its personal Creator). See further Gary W. Deddo, “The Importance of the Personal in the Onto-relational Theology of Thomas F. Torrance,” in T&T Clark Handbook of Thomas F. Torrance, ed. Paul D. Molnar and Myk Habets (London: T&T Clark, 2020), 143-60. 9 T. F. Torrance, Theological Science (Edinburgh: T&T Clark, 1996), xix. 10 Torrance’s approach to the history of theology is largely a matter of drawing connections between theologians he likes (and dislikes). See further Jason R. Radcliff, “Thomas F. Torrance: Historian of Dogma,” in T&T Clark Handbook of Thomas F. Torrance, 101-10, esp. 108. 6

T. F. Torrance (1913–2007)

125

to forms of human experience and conceptualization. Poetics also gives us a precious handle with which to understand Torrance’s antipathy to “analytical, deductive, discursive and linear forms of thought [that] tend to break up the dynamic interrelationality of reality.”11 A kataphatic thinker does not “murder to dissect” (Wordsworth) but rather waits for things to show themselves on their own terms. Dualistic villains: Early modern physics and philosophy. Every good story has a villain, and for Torrance that role is played by dualism in its various guises—epistemological, ontological, and theological.12 The underlying problem of dualism is its division of reality into two incompatible dimensions (e.g., God vs. world, mind vs. body). Dualism shuts God out from having any direct interaction with the world and so contradicts the fundamental Christian axiom that in the humanity of Jesus Christ God has come to us, making it possible that we come to know God. In brief, dualists sever the deep onto-relation between the way something is and the way it is known or experienced. Three brief examples must suffice. René Descartes (1596–1650): The dualism of Western metaphysics. We begin with the father of modern philosophy, René Descartes, a purveyor of his fateful namesake “Cartesian dualism”—the idea that the material body and the immaterial mind are two completely different kinds of substance. For some, Cartesian dualism remains a contemporary option. It is also as old as Plato. Moreover, this metaphysical mind/body dualism helped to give rise to the equally fateful epistemological dualism of subject/object (i.e., the knower and the known), which calls into question the correspondence of what we think we know and what is actually out there.13 Descartes locates rationality in the knowing subject, whereas Torrance insists that true science respects its object’s proper rationality. Colyer, How to Read T. F. Torrance, 17. See his definition in the theological lexicon at the end of T. F. Torrance, ed., Belief in Science and in Christian Life: The Relevance of Michael Polanyi’s Thought for Christian Faith and Life (Edinburgh: Handsel, 1980), 136. 13 Torrance was familiar with James Brown, Subject and Object in Modern Theology (New York: SCM/Macmillan Press, 1955), first delivered as the 1953 Croall Lectures at the University of Edinburgh. 11 12

126

Kevin J. Vanhoozer

Isaac Newton (1642–1727): The dualism of Western physics. We turn next to the father of modern physics, Isaac Newton, who understood space and matter dualistically: “Newton separated space from what happened in it.”14 According to Torrance, he viewed space “as a container independent of what takes place in it”15—namely, the interaction of material particles in motion. Space was absolute: a universal frame of reference in light of which the discrete particles of matter in motion can be plotted with the timeless and universal principles of Euclidean geometry. Newton thus conceived of physical reality as a mechanical universe whose every motion, from falling apples to orbiting planets, is susceptible to mathematical formulation.16 Newton’s achievement nevertheless generated significant problems for both physics and theology. In the first place, Newton never explains what gravity is.17 Theologically the more serious problem was that Newton’s view of the physical universe as “particles in a box” that move according to fixed mathematical laws leaves little room (pun intended) for God to act, much less enter, thus rendering the incarnation of Jesus unthinkable: “If God Himself is the finite Container of all things He can no more become incarnate than a box can become one of the several objects that it contains. Thus Newton found himself in sharp conflict with Nicene theology.”18 Immanuel Kant (1724–1804): The dualism of Western epistemology. Our third dualist, Immanuel Kant, set out to rescue Newton from David Hume’s criticism that, while we may experience one event after another, we do not actually experience the causal relationship on which Newtonian physics depends. Kant’s Copernican revolution proposed an entirely new way of thinking about the relationship of forms of being, expeW. Jim Neidhardt, “Key Themes in Thomas F. Torrance’s Integration of Judeo-Christian Theology and Natural Science,” in T. F. Torrance, The Christian Frame of Mind: Reason, Order, and Openness in Theology and Natural Science (Eugene, OR: Wipf & Stock, 1989), xliii. 15 T. F. Torrance, Space, Time, and Incarnation (Oxford: Oxford University Press, 1969), 22. 16 “Classical physics viewed nature as composed of an infinite number of particles which are connected through external relations.” Colyer, How to Read T. F. Torrance, 56. 17 “Gravity must be caused by an agent acting constantly according to certain laws, but whether this agent be material or immaterial I have left to the consideration of my readers.” Isaac Newton, “Letter to Richard Bentley,” in Newton’s Philosophy of Nature: Selections from His Writings, ed. H. S. Thayer (New York: Hafner, 1953), 172. 18 Torrance, Space, Time, and Incarnation, 39. 14

T. F. Torrance (1913–2007)

127

rience, and knowledge—a Copernican poetics. Space and time for Kant are not things we experience but rather conditions for experience. Causality is not a form of experience but a form of thinking about experience supplied by the human mind (in Kant’s terms, a “category” of the mind). What we can know, and what Newton’s laws of motion describe, are not things-in-themselves but how things in space and time appear to us when processed by the categories of the mind. Science, for Kant, gives us knowledge of phenomena (appearance), not noumena (reality); beyond this limit, reason does not apply.19 Kant represents for Torrance the epitome of the “modern mind,” in all its vainglory, that assigns to itself the privilege of structuring reality. It is a mentality that operates with a poetics of invention, imposing forms of intelligibility onto experience, in contrast to a poetics of discovery that does not create but perceives the intrinsic intelligibility of nature. The social constructivism of many postmoderns is simply a radicalization of Kant’s poetics: both view the mind (human subjectivity) as the power that gives form to reality. As to theology, Kant’s poetics gives rise to a whole series of ugly ditches that reflect the underlying dualism between what is knowable (i.e., what appears in space-time human experience) and what is not, between what the human mind supplies (intelligibility) and what is meaningless in itself (experience). For the modern mind, the incarnation is literally unthinkable: God is by definition that which does not enter into space-time experience, hence the space-time history of Jesus Christ cannot be a revelation of very God, only a noncognitive symbol of morality or expression of piety. What Torrance most regrets about the modern mind, however, is its tendency to dissociate science from theology. Holistic heroes: Nicene theology and Einsteinian relativity. Torrance’s story also includes a number of heroes, all foes of ancient and This was the gist of Kant’s famous attempt to set forth the conditions, and the limits, of human knowledge. See his Critique of Pure Reason, trans. Norman Kemp Smith (London: Macmillan, 1933). Travis Stevick rightly points out the radical difference with Torrance’s approach: “Truly kataphysic knowledge involves probing behind mere appearances to elucidate the source of those appearances. That is to say, it is not enough to give a coherent account of our experience of reality, but we must explain why those appearances are thus and so.” Travis Stevick, “Theological Science Then and Now,” in T&T Clark Handbook of Thomas F. Torrance, 123.

19

128

Kevin J. Vanhoozer

modern dualisms. The kataphysical poetics that these figures champion is what enables Torrance to overcome the fatal divide between ontology and epistemology. Holism here stands for the unitive relation between nature and knowledge. Ancient: Nicene theology (trinitarian relations). Pride of place in Torrance’s roll call of kataphysic faith must go to the fourth-century theologian Athanasius, the chief architect of trinitarian orthodoxy. Torrance’s tireless advocacy of Athanasius’s work at the Council of Nicaea makes Torrance one of the first ressourcement theologians.20 Torrance is adamant that the Greek fathers were aware of the kataphysic nature of true science—namely, that knowledge depends on rightly grasping the inherent nature of the particular reality under investigation. This is what the doctrine of the Trinity is all about: that the Son is the one who makes the Father truly known—and not simply by uttering true statements. What is at stake in the deity of the Son is the claim that God has revealed his true nature to us in physical reality—the body of Jesus—without ceasing to be God and without reducing God to physical reality. According to Torrance, no one grasped this better than Athanasius. His insistence on homoousios—that the Son is “of the same substance” and has the same nature as God—is the guarantee that God has indeed communicated not simply his appearance but his reality, and this within the structures of our spatiotemporal worldly existence: “What Athanasius did was to think through and set out on a scientific basis . . . the Church’s knowledge of the Father and the Son in their mutual relations, and at the same time to think through and work out the biblical account of the economic condescension of the Son in the form of a servant.”21 This, says Torrance, is the way “that Athanasius assimilated the scientific John Webster notes that Torrance was reluctant to speak with his own voice, preferring “to commend a synthesis of what he took to be the classical tradition of the undivided church.” John Webster, “T. F. Torrance, 1913–2007,” International Journal of Systematic Theology 10, no. 4 (2008): 369. 21 Torrance, “Implications of Oikonomia,” 238. Colyer appeals to homoousios as a prime example of Torrance’s “integration of form” in theology: “The homoousion is a faithful exegetical and theological concept that is neither deduced from the biblical text, nor imposed upon it. . . . [It is] an articulation of the basic constitutive relations in the revealing and redemptive activity of God in Jesus Christ.” Elmer Colyer, “A Scientific Theological Method,” in The Promise of Trinitarian Theology: Theologians in Dialogue with T. F. Torrance, ed. Elmer Colyer (Lanham, MD: Rowman & Littlefield, 2001), 224. 20

T. F. Torrance (1913–2007)

129

method that had been developed in Alexandria, namely, rigorous knowledge according to the inherent structure or nature (kata physin) of the realities investigated.”22 The doctrine of the Trinity developed by the Nicene theologians is nothing less than “the proper outcome of scientific engagement with the reality of God, as God is disclosed in Christ.”23 ­Homoousios is shorthand for the kataphatic insight that “it is in Christ that the objective reality of God is intelligibly linked with creaturely and physical forms of thought.”24 Alexandria was also home to the sixth-century physicist and theologian John Philoponus, whom Torrance credits with helping him discern “the powerful heuristic impact of Christian theology upon the foundations and advance of natural science and of physics in particular.”25 It was the Alexandrian teaching about the Word of God in creation “in the beginning” that led Philoponus to criticize Aristotle’s assumption about the eternality of matter.26 That God created light and endowed it with its distinct force led Philoponus to criticize Aristotle’s static notion of light, thereby anticipating twentieth-century physics. In Torrance’s words, “Philoponus realized that Aristotelian logic applied only to static relations or idealized forms, as in Euclidean geometry, and not to real intelligible relations in the actual dynamic world of space and time.”27 Hence Philoponus’s theology influenced his science, and science returned the favor by encouraging him to deepen his understanding of God’s own dynamic reality, in which “divine being and divine act are one.”28 Torrance claims that Philoponus anticipated both Einstein and Barth, which explains why he holds him in high regard as a kataphysical thinker. Karl Barth (1886–1968). Torrance considered Karl Barth the Athanasius of the twentieth century. Whereas Athanasius emphasized Jesus’ oneness with the Father (the holism of God’s being), Barth preserved the oneness T. F. Torrance, Theology in Reconciliation (London: SCM Press, 1965), 241. McGrath, T. F. Torrance, 162. 24 Torrance, Space, Time, and Incarnation, 17. 25 Torrance, Space, Time, and Incarnation, 3. 26 See especially Richard Sorabji, Philoponus and the Rejection of Aristotelian Science (Ithaca, NY: Cornell University Press, 1987). 27 T. F. Torrance, Theological and Natural Science (Eugene, OR: Wipf & Stock, 2002), 10. 28 T. F. Torrance, Theological and Natural Science, 11. 22 23

130

Kevin J. Vanhoozer

between God-to-us and God-in-himself (the holism of God’s beingin-act).29 It was Barth who helped Torrance achieve his fundamental ­kataphysic insight, that the scientific status of an inquiry is determined primarily by its subject matter, which dictates the method by which an object comes to be known: “We know things in accordance with their natures . . . and so we let the nature of what we know determine for us the content and form of our knowledge.”30 Barth therefore helped Torrance appreciate theology as a distinct science with its own distinct object: God’s self-revelation in Jesus Christ. Barth famously exposited divine revelation in trinitarian terms: God (the Father) reveals himself (the Son) through himself (the Spirit). On the other hand, as McGrath notes, Torrance was later disappointed by Barth’s reluctance to make connections with other disciplines—above all, with the natural sciences.31 Barth wanted nothing to do with natural theology, for he believed that God was rightly known on the basis of his revelation in Jesus Christ alone, not on the basis of making inferences from creation. Yet Torrance, while agreeing that the natural order is not an independent source of revelation, nevertheless held that nature can, from the standpoint of faith, be seen for what it is—namely, something God has created. As we shall see, this opened up possibilities for Torrance to pursue theological science beyond Barth. Modern: Einsteinian physics. We turn now to consider two modern heroes in Torrance’s honor roll of kataphysics, now on the side of the natural sciences: James Clerk Maxwell (1831–1879) and Albert Einstein (1879–1955). In both cases, Torrance finds examples of scientists—­ kataphysical kindred spirits, one might say—whose theoretical reflections are bound up with an intuitive discovery of the structure of the natural world. James Clerk Maxwell changed the face of modern physics. He discovered the speed of light and rejected Newton’s mechanical “Barth essentially applies the Nicene homoousios to epistemology, Christology, the doctrine of the Trinity, and soteriology, creatively extending Athanasius’s own use of the theological concept.” Radcliff, “Torrance: Historian of Dogma,” 110. 30 T. F. Torrance, The Ground and Grammar of Theology (Charlottesville: University of Virginia Press, 1980), 8. 31 Alister McGrath, “Thomas F. Torrance and the Search for a Viable Natural Theology: Some Personal Reflections,” Participatio 1 (2009): 69. 29

T. F. Torrance (1913–2007)

131

model of the universe, replacing it with something more fluid and relational. Torrance suggests that it was Maxwell’s faith, particularly his union with Christ, that allowed him to gain “an intuitive appreciation of the relation of God to his creation” and eventually a feel for the relations inherent in nature. This called for a kind of holistic thinking in which “real dynamic relations have their full value, without being mauled by abstract Aristotelian logic or flattened by Euclidean geometric patterns.”32 Maxwell found himself unable to explain the behavior of electricity, magnetism, and light in mechanistic terms only. Particles seemed related not merely externally but internally; that is, they are “interconnected within dynamic fields of force where the interrelations between the particles are part of what they are.”33 In other words, the relationships between particles are “an intrinsic part of what particles really are.”34 Maxwell became convinced, as a scientist willing to have his theories questioned by the very matter he was trying to understand, that “relation is the most important thing to know.”35 This was a radical new insight in an atomic age inclined to think about little particles the way we do billiard balls. When Maxwell came up with his equations explaining the behavior of electromagnetic particles, he was describing not “external” relations between essentially unrelated particles but “internal” relations that constitute a continuous dynamic field. Torrance finds here a compelling parallel with Nicaea’s emphasis on the field of relations or interrelations between the persons of the Trinity (“onto-relations”).36 If Einstein did not exist, Torrance would have to invent him, for Einstein provides Torrance with perhaps his most important kataphysical counterpart in the natural sciences.37 Einstein overturned Newton’s Torrance, Theological and Natural Science, 13. Colyer, How to Read T. F. Torrance, 56n5. 34 Neidhardt, “Key Themes,” xxix. 35 Cited in Torrance, Theological and Natural Science, 14. Einstein held Maxwell’s equations to be the most important event in physics since Newton, and Torrance himself likens them to John Philoponus’s revolutionary transformation of Aristotelian science. Torrance, Theological and Natural Science, 15. 36 T. F. Torrance, Transformation and Convergence in the Frame of Knowledge (Grand Rapids, MI: Eerdmans, 1984), 230. 37 Neidhardt sees seven important continuities between Torrance’s theology and Einstein’s relativity theory. Neidhardt, “Key Themes.” See also Christopher B. Kaiser, “Humanity in an Intelligible Cosmos: Non-Duality in Albert Einstein and Thomas Torrance,” in Colyer, Promise of Trinitarian Theology, 239-67. 32 33

132

Kevin J. Vanhoozer

dualism of space and time, objecting to the imposition of Euclidean geometry (a theoretical system of necessary relations independent of time and space) onto physical reality (that which appears to the senses). As Torrance notes, “Euclidean geometry is suitable for flat spaces but not for curved ones.”38 Einstein’s genius was to call for a new kind of geometry that would be kataphysical—namely, in accord with real (i.e., dynamic) physical reality, understood now not in terms of independent matter in motion in a spatial vacuum that was only externally related but in terms of elements in an internally related space-time gravitational field. Einstein derived his theory of relativity not from prior ideas or from bare sense observations but from an imaginative insight into the objective intelligibility of the physical universe. This, for Torrance, is true objectivity: to let reality object to our theories when those theories seek to force it into procrustean beds.39 Einstein gained objective knowledge of “an inherent relatedness that characterizes the universe independently of our perceiving and conceiving of it.”40 Torrance calls this breathtaking discovery the homoousion of physics—namely, “the basic insight that our knowledge of the universe is not cut short at appearances or what we can deduce from them, but is a grasping of reality in its ontological depth.”41 Torrance considers Einstein a realist: one who holds that “the simplest explanation of what makes theories work is that they relate to the way things really are.”42 Although Torrance believes that Einstein’s theory corresponds to the way things are, there is no “logical bridge” that allows us to read our concepts off of experience directly.43 On the contrary, the invisible (e.g., gravitational forces) explains the visible (e.g., the movement of planets). The wonder of science is “the openness of the structure of the universe to our rational investigation, and the openness of our knowing Torrance, Theological and Natural Science, 41. For more on Torrance and objectivity, see Stevick, “Theological Science Then and Now,” 115-18. 40 Torrance, Ground and Grammar of Theology, 162. 41 Torrance, Ground and Grammar of Theology, 162. 42 McGrath, T. F. Torrance, 215. 43 T. F. Torrance, Reality and Scientific Theology (Edinburgh: Scottish Academic Press, 1985), 76. Cf. Einstein: “There is no inductive method which could lead to the fundamental concepts of physics.” Albert Einstein, “Physics and Reality,” in Out of My Later Years (New York: Philosophical Library, 1956), 83. 38 39

T. F. Torrance (1913–2007)

133

to the intelligible nature of the universe.”44 In Einstein’s words, “The world of our sense experience is comprehensible. The fact that it is comprehensible is a miracle.”45 Theologians, start your engines! How did Einstein come up with his theory of relativity, if not by induction or deduction? Was he simply a better listener to the music of the quantum spheres? The answer, again, is kataphysis. To repeat: it has everything to do with remembering that the truly scientific approach lets the object of inquiry direct the inquiry. Torrance maintains that the physical universe disclosed itself to Einstein’s mind. While this might seem far-fetched, Torrance argues that this is how science progresses from one paradigm to another, through intuition and insight, followed by testing. The true natural scientist, says Torrance, relies on “intuition”— namely, “the sheer weight or impress of external reality upon his apprehension.”46 Some might call it revelation. In a stunning allusion to Moses in Exodus 33:23, Torrance writes, “nature manifests itself to us and even discloses to us objective structures that are inherently non-observable but constitute, so to speak, the invariant back-side of reality.”47 Space-Time Creation and Incarnation: Hommage to Homoousios To this point we have seen how Torrance views both the natural sciences and theology as kataphysical inasmuch as each attempts to think its respective objects according to their own natures. Yet there is also overlap: both theological and natural science have a common interest in the natural world. The incarnation—the fact that the Creator has entered into creation (“The Word became flesh and made his dwelling among us” Jn 1:14)—requires this, and the Nicene homoousion is the church’s effort to state this truth. This overlap threatens Stephen Jay Gould’s idea of nonoverlapping magisteria—that is, the notion that science and religion are masters in their different realms.48 How does Torrance negotiate this Torrance, Reality and Scientific Theology, 53. Einstein, “Physics and Reality,” 65. 46 Torrance, Theological Science, 118. 47 Torrance, Reality and Scientific Theology, 147. 48 Stephen Jay Gould, Leonardo’s Mountain of Clams and the Diet of Worms: Essays on Natural History (New York: Harmony Books, 1998), 269-84. 44 45

134

Kevin J. Vanhoozer

potential conflict of interests? He seldom gets tied up in the usual debates about creation (e.g., about old or young earth creationism or about the length of the days). Instead, he gives most of his attention to the that and the what of the universe: that it did not have to be, and the order it happens to have. The contingency, coherence, and capacities of the created universe are ultimately explicable, he believes, in light of Christology. The contingency of creation: Existence. Matter is not eternal, but the triune God is. God has always been Father, Son, and Spirit, but God has not always been “maker of heaven and earth.” The universe was not a necessary emanation of God’s being. On the contrary, God in love and freedom decided to create the universe ex nihilo. It follows that the universe is neither self-starting, self-sustaining, nor self-explanatory, but rather wholly dependent on God. This is why Torrance insists that science will never be able to provide a sufficient explanation as to why there is something rather than nothing. There is a limit to what the natural sciences can tell us about the universe. The coherence of creation: Order. God’s freedom in creating pertains not only to the fact that he created but also what he created. This is the second sense in which the physical universe is contingent. Although creation has an integrity of its own, it could have been different (for to say otherwise is to deny God’s freedom in creating). It is precisely because the order of the universe cannot be deduced from something like the idea of “best possible world,” or any other a priori, that we have to study it in all its particularity. Not only the existence but also the intelligibility or order of the universe is contingent. Torrance believes this insight to provide the natural sciences with their own Magna Carta: “Belief in order, the conviction that . . . reality is finally and intrinsically orderly, thus constitutes an ultimate regulating factor in all rational and scientific activity.”49 At the same time, because “through him all things were made” (Jn 1:3), the Word who was in the beginning with God (Jn 1:1-2), there can be only one Logos, only one principle of intelligibility in the universe. Torrance T. F. Torrance, “The Transcendental Role of Wisdom in Science,” in Science et sagesse: Entretiens de l’Académie Internationale de Philosophie des Sciences, 1990, ed. Evandro Agazzi (Fribourg: Éditions Universitaires Fribourg Suisse, 1991), 67.

49

T. F. Torrance (1913–2007)

135

views creation as having been inscribed with christological coherence: “For in him all things were created . . . and in him all things hold together” (Col 1:16-17). The universe is “a cosmic unity due to the all-embracing and integrating activity of the divine Logos, so that a single rational order pervades all created existence contingent upon the transcendent rationality of God.”50 This conviction entered the bloodstream of Western intellectual history thanks to the Nicene vision of creation ex nihilo, although various dualisms repressed it for a millennium, thus slowing the development of the physical sciences. It is precisely creation ex nihilo— the idea that God freely created the universe from nothing—that gives science the confidence that there is an underlying order to be discovered and made intelligible. The Nicene emphasis on incarnation implies something else of great significance about the created order—namely, that it is not closed in on itself (as Newtonian physics assumes) but open for divine business (i.e., receptive to divine action). The homoousion explodes the “container” concept of space by informing us that the incarnate Son shared our earthly space while remaining “very God of very God”: “The Nicene alternative conceived of space dynamically, as determined by that which occupies it.”51 Space cannot “contain” the Son; he is Lord of space. He does not simply exist at a particular position, he occupies it: “Space is here a predicate of the Occupant, determined by his agency.”52 The Son is the “place” where the Father makes himself known, engaging the world in self-communicative action. Theological science and modern physics agree: space is not a place for two entities to collide but rather the ground of their interrelatedness—their “field.” The incarnate Christ shares our bodily space-time existence yet is eternally “in” the Father through his onto-relation (i.e., oneness) as his Son. Accordingly, “It is in Christ that the objective reality of God is intelligibly linked with creaturely and physical forms of thought.”53 Kataphatic theology and kataphatic science here agree in thinking of space as a field of Torrance, Theological and Natural Science, 36. P. Mark Achtemeier, “Natural Science and Christian Faith,” in Colyer, Promise of Trinitarian Theology, 280. 52 Torrance, Space, Time, and Incarnation, 15. 53 Torrance, Space, Time, and Incarnation, 17. 50 51

136

Kevin J. Vanhoozer

dynamic relations, although theology sees a vertical as well as horizontal relation to God. The point, not to be missed, is that if we take our bearings from the incarnation, we see that space and time “are open to the transcendent ground of the order they bear within their nature.”54 The competence of creation: Freedom. In addition to its divine and contingent order, Torrance also posits a contingent freedom of creation. Created entities have competence; that is, they are capable of acting according to their natures; they can be themselves. God and creation are not in a competitive relation.55 Yet Torrance also insists, following Scripture, that our world has fallen into disorder as a consequence of human incompetence—that is, their rebellious and irrational abuse of freedom. Homoousios here comes into its own, for all things were not only made through the Logos but are healed and restored in him. The salient point is that the incarnation is not an “interruption” of the natural order but “the freely chosen way of God’s rational love in the fulfillment of his eternal purpose for the universe.”56 Building (In)Dwelling Thinking Nature: Believing Thomas As I mentioned at the outset, Torrance never backs down when reality is the topic, which is one reason he takes the theology/science dialogue so seriously. But what kind of realist? He is not a metaphysical realist, if by metaphysics we mean the belief that unaided human reason can pierce to the division of the joints and marrow of reality, including the relation between Creator and creation. Nor is he a critical realist in the Kantian sense who believes that reason can never pierce deeper than the phenomena of sense experience.57 On the contrary, Torrance is a kataphysical realist who holds that humans can know the reality behind the surface Torrance, Space, Time, and Incarnation, 18. A point made by several theologians across the Christian tradition and perhaps most pointedly in recent years by Kathryn Tanner, God and Creation in Christian Theology (Minneapolis: Fortress Press, 1988). 56 T. F. Torrance, Divine and Contingent Order (Edinburgh: T&T Clark, 1981), 24. See also Rowan Williams, Christ the Heart of Creation (London: Bloomsbury Continuum, 2018). 57 Torrance does call himself a “critical realist” but in a kataphysical (not Kantian) sense inasmuch as he is radically committed “to reality rather than any particular explanation or articulation of that reality.” Stevick, “Theological Science Then and Now,” 120. 54 55

T. F. Torrance (1913–2007)

137

of physical appearances, but only by being open to reality’s self-communication: “In his view, there is no genuinely scientific thinking without this realist attitude in which it is demanded that human knowledge strictly conform to the nature (kata phusin) of the object known.”58 Torrance’s kataphysical approach discerns a fundamental harmony between world and mind, between forms of being (the orderly/intelligible structure of real things) and forms of knowing and speaking (their conceptual representation), a harmony established by creation ex nihilo and instantiated by the incarnation of the Son, who is homoousios with the Creator (and with the human creature). It is to preserve the integrity of this deep, primordial harmony that Torrance appeals to the homoousion—a doctrine that challenges any dualisms that divide what we can know (appearances) from the way things are (reality). I welcome Torrance’s emphasis on theological realism and his insistence on the ­multilevel nature of reality. Nevertheless, I continue to wrestle with certain aspects of his thought. A tale of two Thomases: Toward what kind of natural theology? My first set of questions concerns Torrance’s vision for the partnership of theology and the natural science and his “new” natural theology. How does his revised natural theology relate to the traditional natural theology of Thomas Aquinas, scholasticism in general, and to the Reformed idea of the “book of nature”? In particular, is the scholastic tradition guilty of dualism, as Torrance’s story of Western intellectual history suggests? Must Thomas (Torrance) doubt Thomas (Aquinas)? Relatedly, what can modern science contribute to natural theology (or a theology of nature)? In particular, what authority, if any, does Torrance accord modern science’s reading of nature over that of theology? Thomas Aquinas’s scholasticism: Traditional natural theology. How ought we account for Torrance’s antipathy to his medieval namesake, Saint Thomas, despite the latter’s orthodoxy and affirmation of theology as a science? Everything depends on the underlying assumption of what science is and how it works. Torrance associates Thomas Aquinas with the metaphysical project of trying to gain knowledge of ultimate reality Tapio Luoma, Incarnation and Physics: Natural Science in the Theology of Thomas F. Torrance (Oxford: Oxford University Press, 2002), 61.

58

138

Kevin J. Vanhoozer

(God) on the basis of penultimate reality (nature). Natural theology is the attempt to build a Tower of Babel in which human beings attain heaven, the knowledge of God, from below. Torrance sees Aquinas as complicit in a kind of dualism that seeks to build a “logical bridge” between earth and heaven by deducing knowledge of the supernatural (God) from knowledge of the natural (world). Torrance is reluctant to pursue theological science from below, which is what he sees Aquinas doing. He shares Barth’s low opinion of the “analogy of being”: what God is like cannot be determined by studying nature alone—that is, in abstraction from God’s definitive revelation in Jesus Christ. Scholasticism leads to a view of God that is analogous to nature, yes, but not to God’s nature. It gets us as far as classical theism but not Christian (i.e., trinitarian) theism. We need the homoousion for that. To seek knowledge of God anywhere but the incarnation is “to contradict the Trinity and to set aside the gospel.”59 To the ambition of traditional natural theology, Torrance therefore adds his Scottish “nae!” to Barth’s “nein!”60 Thomas Torrance’s post-Barthian turn: Toward a revised Reformed natural theology. Given his belief in the contingent order of nature, Torrance denies any necessary connection between created order and Creator such that one could draw straight lines from features of the universe to divine perfections: “Torrance characteristically makes this point by saying there is no logical bridge for moving between the created order and God.”61 So far, so Barthian. Torrance is not content to leave it at that, however, for he does not want to leave open the possibility that there is a contrast, much less a dualism, between natural and revealed theology. Rather, he wants to reform natural theology by relocating it within revealed theology, similar to the way in which Einstein reformed Euclidean geometry by locating it within relativity theory.62 So, for Torrance, do the heavens declare the T. F. Torrance, The Christian Doctrine of God: One Being Three Persons (Edinburgh: T&T Clark, 1996), 24. 60 See Karl Barth and Emil Brunner, Natural Theology: Comprising “Nature and Grace” by Professor Emil Brunner and the Reply “No!” by Dr. Karl Barth (London: Geoffrey Bles, 1946). 61 Achtemeier, “Natural Science and Christian Faith,” 276. 62 According to Alister McGrath, “Torrance felt that Barth’s failure to engage with the natural sciences constituted perhaps his most serious weakness.” McGrath, T. F. Torrance, 198. 59

T. F. Torrance (1913–2007)

139

glory of God? Yes, but again, this is not something we can “read off ” of the surface of creation, because there is no necessary or logical relation between God and the world (remember contingency). On the contrary, this is something we must read through the lens of Scripture with the illumination of the Spirit, as Calvin said. Of course, the natural sciences do not look at nature through Scripture. What Torrance insists on, and which Calvin did not, is that theology shares with the natural sciences, formally, the privilege and responsibility of thinking about the world kata physis (i.e., according to its nature) and, materially, the affirmation of the universe’s contingent order and freedom. These are the common commitments that give Torrance hope for a productive dialogue between theology and the natural sciences. At the same time, because the natural sciences cannot in and of themselves yield knowledge of the triune God (only revelation in Christ can do that), some have suggested that Torrance would be better off speaking not of a “new” natural theology but rather of a theology of nature.63 In any case, Torrance’s revised natural theology prompts three questions. First, although the created order is contingent (which for Torrance means it could have been otherwise), how different, how much other, could it have been? Take, for example, the human creature, whom God created in his own image and likeness (Gen 1:26). How much different could humans be and still be in the image of God? Similarly, “the heavens declare the glory of God” (Ps 19:1). Would any astronomical arrangement do, or is there a divine grammar to which these celestial statements must conform? Of course, the existence of the universe is contingent, in that God did not have to create it, but, having decided to create it, how different could things be if all things that are made are made through Christ (Jn 1:3)? Would every possible world display the same fundamental ontorelatedness, for example? Second, do Einstein’s concepts have a better purchase on ultimate reality than Aristotle’s? Or does such a question amount to comparing Newton’s apples (physics) and oranges (metaphysics)? To be sure, Einstein’s See Myk Habets, Theology in Transposition: A Constructive Appraisal of T. F. Torrance (Minneapolis: Fortress Press, 2013), 85-91; and Paul D. Molnar, “Natural Theology Revisited: A Comparison of T. F. Torrance and Karl Barth,” Zeitschrift für dialektische Theologie 21 (2005): 53-83.

63

140

Kevin J. Vanhoozer

dynamic and relational conception of reality explains more about our physical universe than Newtonian physics, but what fascinates Torrance is its parallel to the dynamic relational reality of the triune God. Is there not a danger of falling into the same trap as traditional natural theology— namely, of appealing to the way the world is to shore up one’s doctrine of God? At the very least, Torrance appeals to contemporary natural science as a confirmation that his critique of dualism is correct. Moreover, does not Aquinas do what Christian theologians always do— namely, appropriate the concepts of the contemporary culture in order to reflect biblical judgments (and, if necessary, amend them, as he does with Aristotle’s notion of the eternality of matter)? Matthew Levering is a present-day theologian who claims that we still need Aquinas’s metaphysical categories to interpret Scripture rightly.64 Gerard Verschuuren’s book Aquinas and Modern Science makes a similar point about the necessity of Aquinas’s metaphysics for understanding both classical and quantum physics.65 It may be possible to deploy Aquinas’s metaphysical categories without having to aid and abet his natural theology. My point is this: nature is one, but there may be more than one conceptual scheme for coming to apprehend it. I think Torrance would agree: “When Torrance describes himself as a ‘realist’ . . . he is highlighting his radical commitment to reality rather than any particular explanation or articulation of reality.”66 The purpose of theorizing is not to capture reality in words and concepts but “to facilitate our contact with reality to reach ever more appropriate knowledge according to the reality’s nature.”67 This leads to my final question and to the heart of Torrance’s kataphysical poetics.68 Theological science, personal disclosure, and the conflict of interpretations. How, and under what conditions, can we be sure that our concepts Matthew Levering, Scripture and Metaphysics: Aquinas and the Renewal of Trinitarian Theology (Oxford: Blackwell, 2004). 65 Gerard M. Verschuuren, Aquinas and Modern Science: A New Synthesis of Faith & Reason (Kettering, OH: Angelico Press, 2016). 66 Stevick, “Theological Science Then and Now,” 120. 67 Stevick, “Theological Science Then and Now,” 122. 68 What I am calling kataphysical poetics is similar to what Colyer refers to as “the integration of form” (see his How to Read T. F. Torrance, chap. 9). In both cases, what is at stake is how the knower comes to have concepts (forms of thinking) that correspond to reality (forms of being). 64

T. F. Torrance (1913–2007)

141

and theories are indeed in touch with reality, whether the reality in question is that of God or the natural world? In particular, how can we be sure that we have read the “real text” of Scripture (which Torrance identifies with Christ’s humanity) or the “real text” of the book of nature correctly? Stated differently, How does kataphysics actually work? Torrance insists that reality reveals itself to those with ears to hear and eyes to see, natural scientists and theologians alike. Theories are not simply conceptual projections by a creative mind onto inert matter; they are rather “transparent disclosure models through which . . . the truth in the creation as it has come from God . . . shines through.”69 Torrance’s kataphysical realism maintains that any external reality (the triune God, the physical universe) exerts pressure on our minds as it gives itself to be known, thereby setting up an actual correspondence between our thinking and the way things are—if we as knowers are rightly attuned to what is being given. This is a mighty and momentous if, and mighty difficult to understand, but let me at least make a provisional attempt in the hope that I have heard Torrance aright. Here we can return to Einstein’s imaginative insight beyond the Newtonian surface into the very nature of things, into their ontological depths. Einstein’s theory of relativity is to physics what Athanasius’s doctrine of the deity of the Son is to theology, which is why Torrance refers to Einstein’s breakthrough as “the homoousion of physics”70: the integration of epistemology (forms of knowledge) and ontology (forms of being). Torrance turns to Michael Polanyi to help in interpreting such “aha” moments. There is a tacit dimension to knowing, a subsidiary awareness of background patterns, in addition to focal knowing. Observation and empirical experimentation is not enough: we have to indwell a particular field of investigation in order to grasp its internal constitutive relations. For example, one begins to understand Torrance himself only after indwelling the body of his works—eating, drinking, and dreaming Torrance—for several weeks (I speak from experience). T. F. Torrance, “A Pilgrimage in the School of Christ—An Interview with T. F. Torrance, by I. John Hesselink,” Reformed Review 38, no. 1 (1984): 63. On disclosure models, see further Stevick, “Theological Science Then and Now,” 121-24. 70 Torrance, Ground and Grammar of Theology, 162. 69

142

Kevin J. Vanhoozer

One author thinks the key to understanding Torrance is the notion of compulsion: “Reality compels us to know it according to its true nature.”71 Real things exert real force, yet persons have to be open to it. Polanyi uses perceptual integration as a model. Consider a stereogram from one of the Magic Eye picture books. No amount of detailed analysis of the surface features of the picture enable one to see the three-dimensional image— the “form”—concealed therein. Instead, you have to attend to the whole, without focusing on the details. The mind then “integrates the subsidiary clues to the matrix of intrinsic interrelations between the parts that constitute the three-dimensional whole.”72 How do you know you have seen the image correctly? You will know it when you see it. It is a matter not of forcing your reason on an object but rather letting it “interpret itself as we develop appropriate structures of thought under its impact upon us.”73 As with visual patterns that come together to form a 3D image, so with everything else. We have to be open to the patterns that are pressing in on us from the world. Stated differently, we have to adopt an obedient posture that is receptive of revelation.74 Torrance wants to say something similar about reading Scripture. His Fuller Theological Seminary lectures (published as Reality and Evangelical Theology) criticized the way in which some biblical and theological departments operate with logicalcausal patterns of thought that, given their subject matter, are unscientific, getting no farther than surface phenomena. Exegetes should not simply parse terms and outline grammar but rather prayerfully indwell the text until they perceive (hear, not see) its unitary truth, the living Word of God.75 Torrance calls this “depth-interpretation,” but we could also think of it as kataphysical.76 Revolutionary scientific discoveries are Luoma, Incarnation and Physics, 148. Colyer, How to Read T. F. Torrance, 338. 73 Torrance, Ground and Grammar of Theology, 97. 74 See John C. McDowell, “Thomas F. Torrance on the Doctrine of Revelation,” in T&T Clark Handbook of Thomas F. Torrance, 127-42, esp. 138-40, where McDowell raises questions about how we know when we have sufficiently gotten ourselves out of the way of revelation, as it were, so that we can be sure we have received the object and not our own thought about the object. 75 “It is not just the grammatical connection of the words that give them their meaning but the nature of the reality to which they refer.” T. F. Torrance, Divine Meaning: Studies in Patristic Hermeneutics (Edinburgh: T&T Clark, 1995), 236. 76 See the discussion in John Webster, “T. F. Torrance on Scripture and Hermeneutics,” in The Domain of the Word: Scripture and Theological Reason (London: T&T Clark International, 2012), 86-112, esp. 100. 71 72

T. F. Torrance (1913–2007)

143

eureka moments, not because “I have found it” (contra Archimedes) but because the object of knowledge has found me. Knowing something means opening myself to receive an object’s self-revelation. For example, we hear God’s word in the Bible not when we parse the verbs correctly but when we “listen and repent of what we want to make the Bible say, to listen in such a way as to let the Bible speak against ourselves, that is to listen indeed to the Word of God.”77 I think I understand, but how can I be sure that I have plumbed the depths of Torrance’s thought when others who have also read his work come to a different opinion? How do we account for disagreement? Is there provision for legitimate differences of interpretation? John Hick once told Torrance that he had not understood a word of his lecture, to which an exasperated T. F. replied, “You will never understand, unless you repent!”78 Yet this merely pushes the problem back. Which of the many Christian traditions, all of whom are made up of repentant sinners, is closer to discerning the intrinsic rationality and order of creation and redemption? Which party needs to repent of its poor theological science: Arminians or Calvinists, federal Calvinists or evangelical Calvinists, conservative or progressive evangelicals (no doubt Lutherans would respond, “All of the above”)? There have always been conflicts of interpretations in both theological science and the natural sciences. I am not asking Torrance to make these conflicts go away, but I am asking how he negotiates interpretive disagreements given his kataphysical commitments. What rational recourse do we have if, after repenting and opening ourselves up to God’s Word, we do not all hear the same thing? It is a practical question about how to handle epistemic disagreement. Torrance sees himself as an evangelist, and having a regenerate mind is indeed a prerequisite for Christian theology, but how do we explain, and handle, disagreements over how to understand the natural world, or the Bible, between different Christian traditions? These are large questions, and Torrance is not without resources for responding (e.g., the appeal to catholicity).79 Perhaps it was Thomas F. Torrance, “Introduction: The Place of Christology in Biblical and Dogmatic Theology,” in Essays in Christology for Karl Barth, ed. T. H. L. Parker (London: Lutterworth Press, 1956), 30. 78 Recounted in Farrow, “T. F. Torrance and the Latin Heresy,” 26. 79 Naomi Oreskes raises a similar question, “Why trust science?” and makes a similar appeal to catholicity or, in her terms, consensus. See Naomi Oreskes, Why Trust Science? (Princeton, NJ: Princeton University Press, 2019). 77

144

Kevin J. Vanhoozer

because he knew that disagreement threatened his conception of theological science (kataphysis) that he spent so much time and energy in ecumenical discussion, particularly between East and West.80 Kataphysics and middle-distance realism. My final point concerns the pastoral implications of adopting kataphatic forms of living as well as knowing—that is, forms of life, speech, and thought that accord with how things give themselves to be known. Biblical wisdom is associated with flourishing life because it encourages people to live along the grain of the created order. Kataphysis would thus seem to be a prescription for wisdom, not only knowledge. Torrance rightly acknowledges that Scripture depicts God freely engaging the contingent order of nature he has created. Scripture does not describe the natural world as Maxwell and Einstein do. To what extent, then, can God-talk be coordinated with, or take its cue from, say, quantum mechanics? Jesus says, “Two men will be in the field; one will be taken and the other left” (Mt 24:40), but it is doubtful he had Heisenberg’s indeterminacy principle in mind. The salient point is that Jesus’ parables, like the rest of Scripture, focus on the “mezzo” level, between the micro and the macro. The “mezzo” level corresponds to “middle-distance realism,” a concept I take from David Ford, who in turn took it from the literary critic J. P. Stern’s work On Realism.81 Realism here is a literary category and refers to a level of reality poised somewhere between the quantum and the astronomical. Ford explains, “The middle distance is that focus which best does justice to the ordinary social world of people in interaction.”82 Torrance initiated a dialogue between Reformed and Orthodox churches in which he proposed a solution to the long-standing debate over the Filioque clause that separated Western and Eastern formulations of the doctrine of the Trinity. See further Matthew Baker and Todd Speidell, eds., T. F. Torrance and Eastern Orthodox: Theology in Reconciliation (Eugene, OR: Wipf & Stock, 2015). 81 See David Ford, “System, Story, Performance: A Proposal About the Role of Narrative in Christian Systematic Theology,” in Why Narrative? Readings in Narrative Theology, ed. Stanley Hauerwas and L. Gregory Jones (Grand Rapids, MI: Eerdmans, 1989), 191-215; and J. P. Stern, Realism (London: Routledge/Kegan Paul, 1973). See also Robert C. Greer, “Post-Foundational Middle-Distance Realism,” in Mapping Postmodernism: A Survey of Christian Options (Downers Grove, IL: InterVarsity Press, 2003). Note, however, that Greer uses the term idiosyncratically, as a label for the middle position between metaphysical realism and anti-realism. 82 Ford, “Story, System, Performance,” 195. 80

T. F. Torrance (1913–2007)

145

Newton’s laws still apply to the world humans typically experience. While the salvation we have in Christ is cosmic in scope, Scripture focuses on how our new life in Christ gets worked out in ordinary situations. Moving either too close or too far away results in a loss of the narrative form and content. How can we integrate middle-distance realism, and the narrative form, into Torrance’s theological science and new natural theology? Torrance would no doubt resonate with the concern to do justice to the level of the interpersonal, yet he tends to dialogue much more with the natural rather than the human sciences. By way of contrast, Paul Ricoeur argues that narrative is precisely the literary (and cognitive) form that best accords with human temporality (i.e., being in time).83 Is there room for the middle distance in Torrance’s realism inn? Stated differently, would Torrance acknowledge narrative as a bona fide kataphysical form? He could and should do so, not least because, as Barth reminds us, we think Jesus Christ kata physin precisely by attending to his human history.84 Conclusion: Toward a Glorified Field Theory (and the Poetics of Discipleship) Torrance has given us good reason to agree that Christian theologians should never back down when the topic is reality. What about when the conversation turns to the theory of everything? Einstein sought in vain for a unified field theory that could reconcile general relativity with quantum mechanics. More recently, string theory is all the rage but, appropriately enough, there are still loose ends. Torrance is after a grand—or rather glorified—field theory in which theology and science point together to a larger, unitary intelligibility in which not string but offspring is the operative clue and concept. The Son is the begotten of the Father and the offspring of Abraham (Gal 3:16). All things in heaven and earth—the nature of deity and humanity, the future of the cosmos—are summed up in Christ (Eph 1:10). Christ is the ground and grammar of the shared intelligibility of theology and science.85 For Ricoeur on narrative, see my Biblical Narrative in the Philosophy of Paul Ricoeur: An Essay on Hermeneutics and Theology (Cambridge: Cambridge University Press, 1990). 84 See Karl Barth, The Humanity of God (Louisville, KY: Westminster John Knox, 1996), 37-68. 85 Neidhardt, “Key Themes,” xxxv. 83

146

Kevin J. Vanhoozer

It is therefore fitting to conclude by considering the role of human persons in theological science and scientific theology. Human beings are constituent elements in the natural world. Man is a microcosm of the universe—Homo microcosmos, a roughly six-foot slice of the multiple levels of things that make up reality. Even more remarkable: humanity is that unique place in “the universe whereby it reaches knowledge of itself.”86 We can go further still: humanity lives on the boundary between two dimensions, the physical and the spiritual, at a level of reality which, like all levels, is open to what is above (God) but ultimately inexplicable from below (nature), “so the universe through man at its highest and most advanced level . . . is finally to be understood from its contingent relation to God.”87 Thanks to humanity, the intelligibility of nature and the gospel of Jesus Christ may be voiced. Torrance describes humanity as priests of creation.88 Torrance may have a kindred spirit in Norman Wirzba, who says that we are “in the midst of a crisis of seeing”89 and that “to see the world as God’s creation people must become creatures who live in Christ.”90 Torrance’s kataphysical poetics was perhaps made for such a time as this inasmuch as he reminds us that creation is not simply about origins but the present character of the world and our own place in it. Ultimately, seeing the natural world for what it truly is demands more than theory. It demands cultivating an ethos or form of life that sees the natural world as God’s good creation. To live kata physin, along the grain of the created Torrance, Christian Frame of Mind, 41. Cf. Colyer: “Humanity is thus the one place in space and time where knowledge of sovereign Creator and also knowledge of the contingent creation come to articulation.” Reading T. F. Torrance, 180-81. 87 Torrance, Christian Frame of Mind, 61. 88 Torrance, Christian Frame of Mind, 59-63. See also Eric G. Flett, “Priests of Creation, Mediators of Order: The Human Person as Cultural Being in Thomas F. Torrance’s Theological Anthropology,” Scottish Journal of Theology 58, no. 2 (2005): 161-83. Christ, of course, is the high priest of creation, “through whose service . . . the marvelous rationality, symmetry, harmony, and beauty of God’s creation are . . . given expression in such a way that the whole universe is found to be a glorious hymn to the Creator.” T. F. Torrance, “The Goodness and Dignity of Man in the Christian Tradition,” in Christ in Our Place: The Humanity of God in Christ and the Reconciliation of the World. Essays Presented to Professor James B. Torrance, ed. Trevor Hart and Daniel P. Thimell (Allison Park, PA: Pickwick, 1989), 387. 89 Norman Wirzba, “On Learning to See a Fallen and Flourishing Creation: Alternate Ways of Looking at the World,” in Evolution and the Fall, ed. William T. Cavanaugh and James K. A. Smith (Grand Rapids, MI: Eerdmans, 2017), 162. 90 Wirzba, “On Learning to See,” 173. 86

T. F. Torrance (1913–2007)

147

order, is the way of wisdom that leads to human flourishing. It is a way that coheres with the way of Jesus Christ. Theological science is ultimately a form of discipleship: of listening for the Logos through whom and for whom all things were made, of following the Logos where he leads, of living, materially and in space-time, to the glory of the eternal God. Recommended Reading Primary Works Torrance, T. F. The Christian Frame of Mind: Reason, Order, and Openness in Theology and Natural Science. Eugene, OR: Wipf & Stock, 1989. ———. Divine and Contingent Order. Edinburgh: T&T Clark, 1981. ———. The Ground and Grammar of Theology. Charlottesville: University of Virginia Press, 1980. See chapter three, “Creation and Science.” ———. Space, Time, and Incarnation. Oxford: Oxford University Press, 1969. ———. Theological and Natural Science. Eugene, OR: Wipf & Stock, 2002.

Secondary Works Colyer, Elmer M. How to Read T. F. Torrance: Understanding His Trinitarian & Scientific Theology. Downers Grove, IL: InterVarsity Press, 2001. Colyer, Elmer M., ed. The Promise of Trinitarian Theology: Theologians in Dialogue with T. F. Torrance. Oxford: Rowman & Littlefield, 2001. Habets, Myk. Theology in Transposition: A Constructive Appraisal of T. F. Torrance. Minneapolis: Fortress Press, 2013. Luoma, Tapio. Incarnation and Physics: Natural Science in the Theology of Thomas F. Torrance. AAR Academy Series. Oxford: Oxford University Press, 2002. McGrath, Alister E. T. F. Torrance: An Intellectual Biography. New York: T&T Clark International, 1999. Molnar, Paul D., and Myk Habets, eds. T&T Clark Handbook of Thomas F. Torrance. New York: T&T Clark, 2020. Stevick, Travis M. Encountering Reality: T. F. Torrance on Truth and Human Understanding. Minneapolis: Fortress Press, 2016.

7 J Ü R G E N M O LT M A N N ( 192 6 – ) The Environment of Science and Theology St eph en N. Willia m s Jürgen Moltmann, Professor Emeritus of Systematic Theology at the University of Tübingen, Germany, is one of the most prominent theologians of the second half of the twentieth century. He is known for this “theology of hope” as well as his incorporation of liberation theology and ecology into trinitarian theology. Moltmann grew up in a secular German home. In 1944 he was drafted into the German army, and shortly after his first encounters in battle he surrendered to British soldiers. During his years as a prisoner of war (1945–1948), Moltmann had his first substantive exposure to Christianity, and through chaplains, a Bible, and theological teaching, he converted. After the war, Moltmann decided to pursue theological education. He completed his doctoral degree in 1952 and served as a pastor from 1952 to 1957 before he began teaching in 1958. Following faculty posts in Wuppertal and Bonn University, he was appointed professor of systematic theology at the University of Tübingen (1967). He remained there until his retirement (1994).

J

ürgen Moltmann (1926–) came to prominence in the English-speaking theological world with his Theology of Hope, translated in 1967, and The Crucified God, translated in 1974.1 Since then, Moltmann has been one of the internationally leading Protestant theologians, his name often coupled with that of Wolfhart Pannenberg (1928–2014). Most are still associated with his themes of eschatology, the subject of the first book, and the suffering of God, the subject of the second. A third theme is even Jürgen Moltmann, Theology of Hope: On the Grounds and Implications of a Christian Eschatology (London: SCM Press, 1967); and Jürgen Moltmann, The Crucified God: The Cross as the Foundation and Criticism of Christian Theology (London: SCM Press, 1974).

1

Jürgen Moltmann (1926–)

149

more prominent overall: the Trinity, which frames Moltmann’s developed theological work as a whole. A volume on it launched a sixvolume series of studies in dogmatics that began with The Trinity and the Kingdom of God and ended with Experiences in Theology.2 Although often taken to be a Lutheran theologian, Moltmann is, in fact, Reformed, but his influence has extended far beyond the boundaries of the Reformed tradition and has been widely felt, including in evangelical and liberation theology circles. He cannot be aligned with any school of thought and has sought to engage with the whole Christian tradition while forging a creative theology for our times. He has avoided trying to be systematic in the sense of producing a tight system but has sought to explore various theological themes in an orderly way, consistently attending to their social dimensions and implications. His experience as a German prisoner of war in the Second World War attuned him to the existential realities of human suffering, never far from the surface of his thought. An Interest in Science? Does Moltmann have much to contribute to the dialogue between science and theology where the question of creation is concerned? The most prominent “scientist-theologian” in the English-speaking world is probably John Polkinghorne, whose favorite contemporary theologian is Jürgen Moltmann.3 Yet, at one stage, he lamented the “spectacle of a distinguished theologian [i.e., Moltmann] writing over three hundred pages on God in creation with only an occasional and cursory reference to scientific insight.”4 The offending text was Moltmann’s God in Creation, Moltmann’s large-scale treatment of this subject in his series of dogmatic volumes.5 So we might suppose that our question must be answered in Jürgen Moltmann, The Trinity and the Kingdom of God: The Doctrine of God (London: SCM Press, 1981); and Jürgen Moltmann, Experiences in Theology: Ways and Forms of Christian Theology (Minneapolis: Fortress Press, 2000). 3 Fraser Watts and Christopher C. Knight, eds., God and the Scientist: Exploring the Work of John Polkinghorne (Burlington, VT: Ashgate, 2012), 270. Polkinghorne wrote a volume titled Scientists as Theologians: A Comparison of the Writings of Ian Barbour, Arthur Peacocke and John Polkinghorne (London: SPCK, 1996). 4 John Polkinghorne, Science and Creation (London: SPCK, 1988), 2. 5 Jürgen Moltmann, God in Creation: An Ecological Doctrine of Creation (London: SCM Press, 1985). 2

150

Stephen N. Williams

the negative. Later, however, Polkinghorne wrote of Moltmann and science in a different and positive, albeit quite general, vein.6 These two judgments conveniently frame our approach to the general question of Moltmann’s engagement with science before moving on to creation in particular. There is a basis for each judgment, and we turn to them in order, beginning with the critical one. Moltmann described God in Creation as “an attempt at a doctrine of creation that is compatible with the natural sciences.”7 In his preface to it, he tells the reader that he has both “tried to enter into discussion” with scientific forebears such as Isaac Newton and that “it is the doctrine of creation particularly which will work out points of approach for a theological discussion with scientific findings, hypotheses and theories.” These two statements must be read in context. The first statement is contextualized by the author’s aim of producing an ecumenical theology. This goes beyond drawing on Catholicism, Protestantism, and Orthodoxy to an exploration of the connections between Judaism and Christianity. Moltmann’s reference to “scientific forebears” comes in the course of his observations on ecumenical method in tackling theological questions. He alludes to theological forebears as well, but where only one scientific forebear (Isaac Newton) is mentioned, three theological forebears (Augustine, Aquinas, and Calvin) are listed. This proportion is a sign of the internal balance of theological and scientific engagement both in this particular volume and in Moltmann’s authorship as a whole. Moltmann’s second statement is succeeded by the words “I have tried to take account of the upheaval which the sciences are undergoing today wherever they are pursued in awareness of the ecological crisis.” The subtitle of Moltmann’s volume is An Ecological Doctrine of Creation. It is John Polkinghorne, “Moltmann’s Engagement with the Natural Sciences,” in God’s Life in Trinity, ed. M. Volf and M. Welker (Minneapolis: Fortress Press, 2006), 61-70. 7 Jürgen Moltmann, “God’s Kenosis in the Creation and Consummation of the World,” in The Work of Love: Creation as Kenosis, ed. John Polkinghorne (Grand Rapids, MI: Eerdmans, 2001). Graham Buxton picks out Moltmann’s dialogue with scientists as one evidence of his willingness for ­interdisciplinary engagement. See “Moltmann on Creation,” in Jürgen Moltmann and Evangelical Theology: A Critical Engagement, ed. Sung Wook Chung (Eugene, OR: Pickwick, 2012), 40-68; see p. 42 for this comment. Yet he could write his account of Moltmann’s view of creation with scarcely any reference to science. 6

Jürgen Moltmann (1926–)

151

not that ecological and environmental concerns constantly inform the detailed discussion offered in the volume; rather, they give it its steady orientation. Moltmann does not regard this as just one possible orientation among others to a contemporary doctrine of creation: What we call the environmental crisis is not merely a crisis in the natural environment of human beings. It is nothing less than a crisis in human beings themselves . . . a crisis of life on this planet . . . so comprehensive and so irreversible that it can not unjustly be described as apocalyptic. . . . As far as we can judge, it is the beginning of a life and death struggle for creation on this earth.8

Environmental concern provides the context in which we must learn the permanently valid truth that “the inner secret of creation is this indwelling of God,” the all-pervading Spirit, and it is announced from the beginning of Scripture that “the inner secret of the sabbath of creation is God’s rest”.9 So God in Creation is a theological essay in which the Christian doctrine of creation is adumbrated in the context of environmental crisis and where the scientific interest does not feature independently of that context; that is, it is an interest governed by the theological and ecological concerns announced in the title and subtitle of the volume. Whether or not we agree with the first of Polkinghorne’s judgments with reference to this volume probably depends on what we think is reasonable to expect in the way of engagement with science from a theologian exploring Christian doctrine. The opening chapter of God in Creation scarcely gives science a high profile. “Modern sciences, especially nuclear physics and biology” have shown that “analytic thinking” is reductionistic and barren, analytic thinking being the kind of thinking which holds that scientific understanding proceeds by analyzing the object into its minutest constituents.10 Not much more is said. Moltmann Moltmann, God in Creation, xi. In 2001 Moltmann wrote, “It was only slowly, at the beginning of the 1970s, that we became conscious that human history runs its course within the ecological conditions of the earth.” Science and Wisdom (London: SCM Press, 2003), 111. Ecological concern features from the beginning of his discussion of “The Cosmic Christ” in the volume that follows God in Creation in Moltmann’s dogmatic series, The Way of Jesus Christ: Christology in Messianic Dimensions (London: SCM Press, 1990), 274. 9 Moltmann, God in Creation, xii. 10 Moltmann, God in Creation, 2. 8

152

Stephen N. Williams

is thinking of science very broadly here, and if any scientist inspires him, it is the Austrian physicist Fritjof Capra.11 As he proceeds, Moltmann emphasizes the fact that we should reverse the way in which God’s creation has frequently been understood as “nature,” something that can be exploited “in accordance with the laws science has discovered.”12 What does such reversal look like for the natural sciences? Moltmann briefly surveys the way in which the historical formation of the doctrine of creation has been steered by the relationship between theology and science. He concludes that as in the past “the sciences have shown us how to understand creation as nature,” so now “theology must show how nature is to be understood as God’s creation.”13 Yet when Moltmann proceeds to explore the theology of nature, he again touches on science only in its very broad dimensions and considered as a large-scale enterprise. A seminal chapter on the doctrine of God as Creator includes theological reflection on time and space, but the subsequent chapters devoted to time and space (chaps. 5–6) say little from a scientific point of view. In his preface, Moltmann indicates that the chapter “The Space of Creation” is an exemplary case of his failure to take forward the discussion as far as he had wanted to in this volume. The first of Polkinghorne’s two verdicts seems to be justified. It is when discussing evolution that Moltmann shows most interest in scientific detail, but the “detail” is relative to his engagement elsewhere in the volume; from the scientist’s point of view, there is little detail. What, then, accounts for Polkinghorne’s second and positive judgment? Apparently, the passage of time. This latter judgment was issued in 2006, eighteen years after the unfavorable one. In these intervening years, Moltmann had devoted a volume of essays to the subject, Science and Wisdom.14 Prior to its appearance, Moltmann had completed the series However, Moltmann rejects Capra’s “divinization of evolution.” See History and the Triune God: Contributions to Trinitarian Theology (London: SCM Press, 1991), 133. Alongside Capra, he mentions Erich Jantsch here as a New Age scientist and uses Jantsch’s biological terminology in his description of “Cosmic Spirit” in God and Creation, 17: see 323n16. 12 Moltmann, God in Creation, 21. 13 Moltmann, God in Creation, 38. 14 Presumably, Polkinghorne had also come to appreciate more than he previously did the broad scale of Moltmann’s engagement with science even in God in Creation. In return, the footnotes to his 2011 Boyle Lecture show how much Moltmann came to appreciate Polkinghorne’s work: 11

Jürgen Moltmann (1926–)

153

of studies in dogmatics, of which God and Creation was the second volume, and the title of the epilogue in the final volume, Experiences in Theology, signified what was to come in his authorship: “‘The Fear of the Lord Is the Beginning of Wisdom’: Science and Wisdom.”15 Moltmann styled his epilogue as “a meditation on the relationship between theological wisdom and scientific knowledge.” Reflecting in this volume on his early theological formation, Moltmann mentioned, among other things, his participation in the 1960s in developing a collection titled Theory of Open Systems, whose papers were published by the biologist Ernst Ulrich von Weizsäcker.16 When he comes to the epilogue, Moltmann discourses on the meta-themes of knowledge, power over nature, and wisdom. Yet we still have no scientific detail. Two years later he published Science and Wisdom, one-third of whose essays were previously unpublished. The emphasis in it is what we should expect on the basis of the closing remarks in Experiences, although there is obviously more scientific detail than we find there. Moltmann persistently wants to “see the modern sciences, and especially the biosciences, in their economic and social context. And this context is the scientific-technologicalindustrial complex of modern society.”17 Like theology, science needs renewed wisdom in our times. Moltmann moves beyond this generality in connecting theology and science, specifying that both deal with an open future, the future of nature, and God’s creation. So we are back with eschatology, except that now it is not just human history but the whole created order that is unfolding. Science is a discipline that should be geared to the unfolding cosmos and not to static, ahistorical, structural constants in it.18 “Is the World Unfinished? On Interactions Between Science and Theology in the Concepts of Nature, Time and the Future,” Theology 114, no. 6 (2011): 403-13; See footnotes 3, 26, and, indirectly, 13. The lecture begins with Moltmann’s confession that for him, “science is a ‘Paradise Lost,’” the object of his early and ungratified intellectual passion. 15 Moltmann, Experiences, 334-43. 16 Moltmann, Experiences, 8. Ernst was the son of the renowned physicist and philosopher Carl Friedrich von Weizsäcker, to whose History of Nature Moltmann alludes positively in many writings. 17 Moltmann, Science and Wisdom, 26. 18 As a matter of fact, Moltmann observed early in his literary career that “it is only together with the sciences that eschatological faith can arrive at confidence in history.” This remark was made in 1966 in an essay included in his collection Hope and Planning (London: SCM Press, 1971). I follow the retranslation of this sentence in “Theology in the World of the Modern Sciences,” Science and Wisdom, 23.

154

Stephen N. Williams

Moltmann’s essay “Eschatological Perspectives on the Future of the Universe” in Science and Wisdom is a good example of how he engages with science, and it is appropriate to mention it in connection with creation. According to Moltmann both here and throughout his literature, creation takes three forms: creatio ex nihilo (creation out of nothing), creatio continua (continuous creation), and the nova creatio (new creation).19 In this essay he says that, as he is no scientist, he will discuss “the eschatological perspectives which emerge for the future of the universe from Christian theology” rather than address “eschatology from a cosmic perspective.”20 If Christians subscribe to the eschatological belief that the world will be annihilated, the cosmological analogue is that the universe will revert to something like its condition before the Big Bang. Conversely, if Christians have theological reasons for subscribing to the eschatological belief that the world will be transformed and not annihilated, the cosmological analogue in that case might be described in terms of reading the history of the universe as a universal information process. For himself, Moltmann entertains belief in an unfolding world destined to issue in God’s eschatological future. The main scientific principle correlative to this belief is the openness and complexity of systems, and if we expand the second law of thermodynamics into “the theory of open systems,” we give scientific form to our theological conviction that the universe is open and incomplete. What Moltmann does with the second law of thermodynamics in this volume is instructive if our question is how much Moltmann engages with science. In Science and Wisdom, Moltmann follows up with the chapters “The Time of Creation” and “The Space of Creation” that feature in God in Creation. In his earlier chapter on time, Moltmann does not look at the science, only at the theology; history, not nature, is his principal interest. The same is true in the two chapters he devotes to time in Science and Wisdom. They are not what we should anticipate, for his stated aim is to produce a “temporal concept of time” to mediate “between the objective time-measurements of physics and the subjective temporal The three forms, which are found in the fifth chapter of Science and Wisdom, are addressed in this essay below. 20 He thus reversed the terms of the question discussed in a symposium that included this contribution. 19

Jürgen Moltmann (1926–)

155

experiences of eternity,”21 of which the latter is the subject of theological description. Yet aside from reference to the fact that the second law of thermodynamics introduces into physics a concept of irreversible time that overthrows the Newtonian paradigm—something that supports theological conviction about the irreversibly unfolding history of the world, including nature—we find no engagement with physics. The main line of Moltmann’s thought in this respect is succinctly represented in his Boyle Lecture. Because “the Second Law of thermodynamics introduced into physics in principle the concept of irreversible time,” von Weizsäcker’s history of nature and “the experience of history in the Abrahamitic religions” fit nicely into an alignment of past and future with actuality and potentiality. The rule of irreversibility is expressed in the fact that “potentiality becomes reality but reality never again becomes possibility.”22 When it comes to space, things are not quite the same. In God in Creation, Moltmann expresses the desire to deal with God and space “in the light of new scientific conceptions about the space-time continuum.”23 He does not engage with these new conceptions in Science and Wisdom. However, in the earlier work, he shows an interest in Max Jammer’s work Concepts of Space, to which he returns in his essay “God and Space” in the later work. We also shall return to it as we turn to Moltmann’s treatment of creation.24 In the meantime, we have seen that, despite Polkinghorne’s different verdicts, Moltmann does not show much progressive interest in scientific detail. Implicitly, as far as he is concerned, if the theologian shows awareness of the broad scientific picture and reflects both on scientific culture and on those particular features of the natural world that scientifically bear out a theological perspective on nature or creation, little more should be expected. Perhaps Polkinghorne, for one, came to agree. Moltmann, Science and Wisdom, 85. Moltmann, “Is the World Unfinished?,” 408-9. 23 Moltmann, God in Creation, xiv. 24 The way he links his discussion on time and space in God and Creation with his later discussion of them in an eschatological context indicates the importance Moltmann attaches to theological reflection on these themes in connection with creation. See The Coming of God: Christian Eschatology (London: SCM Press, 1996), 261. 21 22

156

Stephen N. Williams

Creation What does Moltmann believe about creation? We noted that he believes in three forms of it: creatio ex nihilo, creatio continua, and nova creatio. I shall concentrate on the first of these because the last two forms embrace providence and eschatology, themes much too big to be treated within the confines of a single chapter. However, it is important to boldly ­underline that all these forms are comprehended in Moltmann’s understanding of creation. Proceeding within our strict limitations, an account of how Moltmann understands creation must start with the title of God in Creation and ask, Who is God in creation? He answers, “By the title ‘God in creation’ I mean God the Holy Spirit. God is ‘the lover of life’ and his Spirit is in all created beings.”25 So is God in Creation a pneumatology—that is, a theology of the Holy Spirit? The volume that succeeds it in Moltmann’s dogmatic series is on Christology, and the one after that, The Spirit of Life, treats pneumatology.26 Given this sequence, we might expect God in Creation, successor to the first volume in his dogmatic series, The Trinity and the Kingdom of God, to be about God the Father. It is not, and this lends added significance to the title of God in Creation, for Moltmann believes that the Christian doctrine of God and creation will not be properly advanced if it is about God the Father Almighty, Creator of heaven and earth. He draws attention to his studied departure from earlier ways of following the main divisions of the Apostles’ Creed.27 From the inception of his dogmatic series, Moltmann is keen to avoid depicting God in terms of transcendent patriarchy. As he does in that first volume on the Trinity, Moltmann steadily works with a contrast in God in Creation. The contrast is between God as a being who relates to his creation so intimately that we can speak of creation as his cosmic home and God as a sovereign ruler who relates to Moltmann, God in Creation, xii. Respectively, The Way of Jesus Christ and The Spirit of Life: A Universal Affirmation (London: SCM Press, 1992). At the time of publication of God in Creation, Moltmann did not expect to produce a volume on pneumatology following the one on Christology (see xv). 27 Moltmann, God in Creation, xii. The opening chapter in History and the Triune God is “‘I Believe in God the Father’: Patriarchal or Non-Patriarchal Talk of God” (Moltmann’s italics), and the second is “The Motherly Father and the Power of His Mercy.” The “matriarchal subculture” in Christianity is a “resource for renewal.” See Elizabeth Moltmann-Wendel and Jürgen Moltmann, Humanity in God (London: SCM Press, 1983), 49-50. 25 26

Jürgen Moltmann (1926–)

157

his creation as subject to object. Moltmann champions the former conception. The latter has led humans to relate to God’s creation in the corresponding way—that is, as subject to object. It is a recipe for ecological disaster. Pivotal in Moltmann’s theology of creation is the sabbath. Indeed, he is out to develop “a sabbath doctrine of creation.”28 The goal of creation is sabbath rest. God the Spirit is the animating principle of his creation: “The patterns and symmetries, the movements and the rhythms, the fields and the material conglomerations of cosmic energy all come into being out of the community, and in the community, of the divine Spirit.”29 This is an ecological replacement for mechanistic thinking about the cosmos. “An ecological doctrine of creation implies a new kind of thinking about God. The centre of this thinking is no longer the distinction between God and the world. The centre is the recognition of the presence of God in the world and the presence of the world in God.”30 The relation between God and creation is what is called perichoretic, a technical term in the history of theology that signifies an inter-penetration or flowing into each other of different entities.31 In a later essay, Moltmann “translates” into five theological theses Paul’s grand conception “of the torment and the hope of God’s Spirit” with reference to creation in Romans 8.32 These manifest the fruit of the Spirit’s immanence, his indwelling, in creation. First, “the immanence of the transcendent divine Spirit is the foundation and driving power for the self-transcendence of all open systems of matter and life, and of all human forms of life in history.” It is “God in creation” that “makes creation a world open to the future.” Second, that same immanence “is also the foundation and driving power of the evolution of life into ever richer and more complex forms and syntheses.” Third, on account of that immanence, “in all things we see the many-faceted expression of the divine Creator.” Fourth, the Spirit “transforms everything that exists Moltmann, God in Creation, 6. Moltmann, God in Creation, 11. 30 Moltmann, God in Creation, 13. 31 Moltmann, “Is the World Unfinished?,” 407. In the history of theology, the term is most familiarly applied to the relationship between the persons of the Trinity, although that has not been its only application. 32 See Jürgen Moltmann, Sun of Righteousness, Arise! God’s Future for Humanity and the Earth (Minneapolis: Fortress Press, 2010), 206-8. 28 29

158

Stephen N. Williams

into self-transcending movements,” which encourages a positive interpretation of “the dissipative, unbalanced systems of life” captured in chaos theory. Finally, the Spirit “makes the world in which we live a spiritual world,” which strains forward to the sabbath of the new earth. Although God in Creation is the principal source for understanding Moltmann’s view of creation, two preceding contributions are noteworthy, each anticipating the development of some of the main lines of his argument there. The first essay is “Creation as an Open System.”33 In line with the persistent eschatological orientation that marks his theology from the outset, Moltmann advances “an eschatological understanding of creation,” which he believes to be in accord with the Old Testament.34 If we maintain a belief in creatio ex nihilo, we must be sure to stipulate that it is a creatio mutabilis—that is, a creation which is not a static order but open to change. This makes for an open history of the world, a history not fixed any more than creation itself is fixed. Moltmann views the created and historical orders as scenes of conflict between the forces of destruction and of salvation so that the history of God (a phrase that will ring strange in some readers’ ears) is “the opening up in time of closed systems.”35 It is not only the present order that is not in stasis, but the same is true of the eschatological order—it too will be found to constitute a dynamically open scene when God comes to dwell with us in his fullness. In this essay, Moltmann does not confine himself to theological reflection sealed off from science. He notes that science describes a world in which complex systems have evolved into increasing openness and potentiality. The cosmos is not determinate. If the eschatological kingdom of glory perfects a cosmos of this kind, we must conceive of the cosmos in terms of “the openness of all finite systems for infinity.”36 Moltmann appealed to quantum physics to support a theological Jürgen Moltmann, The Future of Creation (Philadelphia: Fortress Press, 1979), chap. 8. It was first published in German in a collection two years prior. 34 Moltmann, Future of Creation, 116. For Old Testament support, he cites Ludwig Köhler, claiming, “In OT theology creation is an eschatological concept” (118). 35 Moltmann, Future of Creation, 126. According to Moltmann, the operations of the Trinity in history—what is called God’s “economy”—should not to be conceived of as the external operations of a static, immutable, and timeless being. To take the economic Trinity seriously is to ascribe to God a history. 36 Moltmann, Future of Creation, 127. 33

Jürgen Moltmann (1926–)

159

conviction about the interrelational, as opposed to the subject-object, nature of human connection with the natural order. His account of creation as an open system issues conclusions about freedom, servanthood, and justice, thus exemplifying an important principle that Moltmann consistently adopts throughout his theology; dogmatics and ethics must be conjoined. Where science enters the equation, its findings are used in combination with dogmatics and ethics to advance a theologically comprehensive outlook on whatever issue is under discussion. The second noteworthy contribution prior to God in Creation is a brief but significant discussion, “The Creation of the Father,” in the first volume of Moltmann’s dogmatic contributions to theology, The Trinity and the Kingdom of God. The difficulty with the way in which creation is traditionally viewed, Moltmann maintains, is that it is “without any significance for God himself.”37 By this he means that creation is traditionally conceived as an entirely contingent act of will; it might or might not have happened. To say that it is an act of will does not mean that if God created, he could have created in just any way whatsoever. For the nature of God lies behind his creation, and the creative act is bound to accord with the Creator’s nature. However, God’s nature is traditionally understood to be that of a self-sufficient being. Moltmann proposes an alternative. We ought to view creation as the outflow of an inner-trinitarian love that seeks to express love for its other.38 This does not mean that we should now regard God as a being who is needily dependent on the existence of a world. No, Moltmann retains belief in creation as free divine self-expression. But we must drop the antithesis: either God is free to create and might not have done so or God is bound by some sort of necessity to create. “In God,” Moltmann declares, “necessity and freedom coincide.” “For God it is axiomatic to love freely.”39 From eternity, God is self-communicating love. This is an attempt to deal with a theological difficulty that Moltmann detects in the tradition, but he thinks the tradition also harbors a metaphysical along with a theological difficulty—that is, a broader sort of Moltmann, Trinity and the Kingdom of God, 105. Inner-trinitarian love for the divine other is the love of like for like; its outflow is love for the unlike human other. 39 Moltmann, Trinity and the Kingdom of God, 107. 37 38

160

Stephen N. Williams

difficulty concerning the way we shape our notions of reality. Creation is traditionally viewed as “outward”; there is God, and what he creates is outside himself. But if God is omnipresent, how can there be an “outside” at all? Moltmann believes there cannot be. It follows that, for God to create, self-limitation on his part is necessary. God achieves this by withdrawing himself so that a space is formed for the nihil (nothingness) in which God creates. Creation thus involves two acts. First, God releases the space in which he can create; second, he creates in the space he has released. This is where Moltmann draws on the work of Max Jammer.40 In his Concepts of Space, Jammer described the background, significance, and impact of the thought of Isaac Luria, the sixteenth-century renovator of mystical, kabbalistic Judaism.41 We are introduced here to the notion of the “Zimzum,” “the divine self-concentration,” which constitutes the creation of “space by self-restriction.”42 In The Trinity and the Kingdom of God, Moltmann reproduces Isaac Luria’s mystical doctrine of zimzum, glossing it as ‘“concentration’ or ‘contraction,’ a withdrawal into the self.” Moltmann supports the judgment that this “is the only serious attempt ever made to think through the idea of ‘creation out of nothing’ in a truly theological way.”43 He claims that it should be positively appropriated and put to use in a Christian trinitarian context that specifies the creation of the world through the Son and by the operation of the Spirit. Creation has its root in and follows from the love of the Father for the Son. Through the energies of the Holy Spirit, creation comes to share in divine inner-trinitarian life without collapsing into pantheistic identification with deity. The two works on which I have just drawn set forth ideas that are taken up, contextualized, expanded, and supplemented in God in Creation. Both open systems and zimzum turn out to be important for Moltmann not only here but elsewhere in his authorship. God as Spirit indwells the open system of creation, which is destined for the kingdom of divine glory. The Christian theological tradition overemphasized Max Jammer, Concepts of Space: The History of Theories of Space in Physics, 2nd ed. (Cambridge, MA: Harvard University Press, 1970). 41 Kabbalistic Judaism is a mystical and esoteric tradition that came to light in the Middle Ages. 42 Jammer, Concepts of Space, 48. 43 Moltmann, Trinity and the Kingdom of God, 105. 40

Jürgen Moltmann (1926–)

161

transcendence at the expense of immanence. Our corrective is to incorporate the Jewish (rabbinic and kabbalistic) understanding of the shekinah, God’s dwelling among us, into a trinitarian framework in the service of an emphasis on divine immanence.44 When it comes to God and creation, it is interrelation and interpenetration, not subject and object, that are the theological order of the day. For Moltmann, what creatio ex nihilo basically means is that the cosmos is dependent on the sovereign will of God for its ongoing and not just its original existence. This creation is a divine kenosis. The language of kenosis is borrowed from an expression in Philippians 2, where the self-emptying of God in Christ is described.45 Moltmann proposes a parallel in God’s action in creation. God’s withdrawal and contraction is the parallel in question, making creation a form of kenosis. As he puts it in a later essay, “Kenotic self-surrender is God’s Trinitarian nature, and it is therefore the mark of all his works ‘outwards.’”46 Now this is the framework within which Moltmann’s reflections on space and time in God in Creation take place. “The space of the world corresponds to God’s world-presence, which initiates this space, limits it and interpenetrates it.”47 Moltmann thus fuses a “Christian and Jewish kenotic theology.”48 There is no mistaking the radicalism of creation kenosis. As Moltmann puts it in that same essay, “God is only Almighty where there is nothing.” Creation as kenosis impacts divine omniscience as well. God cannot foresee the content of human decisions, so “he waits for those he has created, and awaits them. He is curious about the path they will take, for they are his future. He learns from them.”49 Moltmann is loath to contrast transcendence with immanence absolutely. See Spirit of Life, 31-38 on divine “immanent transcendence.” Shekinah is expounded (47-51) but not before we are reminded of the kabbalistic MAKOM (43). “The earliest indication of a connection between space and God lies in the use of the term ‘place’ (maḳom) as a name for God in Palestinian Judaism of the first century.” Jammer, Concepts of Space, 28. 45 Moltmann, God in Creation, 88. 46 Moltmann, “God’s Kenosis,” 140-41. For Shekinah and the kenosis of the Spirit, see Spirit of Life (51), and for the unity of Sabbath and Shekinah, see Coming of God (361-67). 47 Moltmann, God in Creation, 157. 48 Moltmann, “God’s Kenosis,” 137. He later makes the further move, on which I shall not elaborate here, of speaking of the divine withdrawal in creation of his uncreated light. See Moltmann, Science and Wisdom, 119. 49 Moltmann, Science and Wisdom, 120. 44

162

Stephen N. Williams

In God in Creation, although Moltmann believes that theological accounts of creation often place undue emphasis on the opening chapter of Genesis and correspondingly neglect an approach to creation governed by the proper christological center of all theology, he does attend to the Genesis account. We find this not only in his allusion to the sabbath but also in his discussion of heaven and earth, which leads him to consider the scientific account of the created order. On the basis of an exegetical survey of the different uses of the word heaven or heavens in the Hebrew Scriptures, Moltmann notes that the singular use of the word heaven is symbolically and conceptually useful in this respect. Creation is an open system. It includes determined features, and the word earth signifies that, but the word heaven signifies the undetermined side. Here we have a theological description of the interplay that science has discovered between the lawful regularity of nature and its genuine openness in terms of history.50 “Creation lives from the continual inflow of the energies of the Spirit of God . . . because heaven is open . . . the world has a future.”51 With this, we are moving onto creatio continua and nova creatio. Moltmann believes that by placing too much emphasis on creatio originalis, we neglect not only the creatio continua, which is scientifically described in evolutionary terms, but also the nova creatio when we view evolution anthropocentrically. Anthropocentricity undersells humanity, which should be understood neither in terms of its origins nor of its development but fundamentally in terms of its eschatological destiny. As Friedrich Oetinger remarks, “Embodiment is the end of all God’s works.”52 It is comprehensive. With it, we arrive at the sabbath feast of God’s creation. Creation and Science Before drawing conclusions on the way Moltmann relates theology and science to his understanding of creation, we should note that much in Moltmann, God in Creation, 163. Moltmann, God in Creation, 183. 52 Karl Barth, with whom Moltmann frequently expresses his disagreement on God and history, remarks that Oetinger’s statement “was a sound if exaggerated expression of very necessary opposition to the flight of the Enlightenment spirit from nature, but it is not really suited for conversion into a dogma.” Church Dogmatics (Edinburgh: T&T Clark, 1975), I/1:134. 50 51

Jürgen Moltmann (1926–)

163

his work will dishearten adherents of the theological tradition. Does he modify it or abandon it? If he does not abandon it, it is at least a radical modification; for example, he does not offer zimzum as just an edifying tweak on tradition.53 His use of Scripture in theology has also drawn very sharp criticism from a leading, very sympathetic commentator, and Moltmann’s response to this criticism is telling: “Taking account of exegetical disciplines, I can develop my own theological relation to the biblical texts . . . I am a theological partner with the texts which I cite, not their exegete.”54 If I do not follow up theological questions arising from Scripture and tradition, it is only because our specific interest in Moltmann is directed to the way he thinks of theology and science in relation to creation. Late in his authorship, after remarking that “only what corresponds to God’s revelation according to Holy Scripture can be read and understood as divine wisdom in nature,” Moltmann balances the statement with the observation that “only that [in the biblical writings] which is compatible with scientific reason or which opens up for science new horizons of interpretation can count as divine revelation.”55 Initially, it may appear that this pair of statements portends promising guidance in our reading of Moltmann on theology and science. However, it is not really so; the latter statement, in particular, is of very limited hermeneutical use in exploring Moltmann’s work. The reader who works through Moltmann’s corpus is unlikely to be aware of the second of these two principles—the one on new horizons of interpretation. We may see it in operation occasionally, particularly if we begin to look out for it deliberately, but it is not usually on the surface. What about the first principle—that which is biblically incompatible with scientific reason cannot count as revelation? It is tacitly, if not explicitly, taken by Moltmann as self-evident. Moltmann assumes what most theologians assume—namely, that science has For germane critical questions on what Moltmann does with space, see Alan J. Torrance, “Creatio Ex Nihilo and the Spatio-Temporal Dimensions, with Specific Reference to Jürgen Moltmann and D. C. Williams,” in The Doctrine of Creation: Essays in Dogmatics, History and Philosophy, ed. Colin E. Gunton (New York: T&T Clark, 1997), esp. 90-93. 54 Jürgen Moltmann, “The Bible, the Exegete and the Theologian: Response to Richard Moltmann,” in God Will Be All in All: The Eschatology of Jürgen Moltmann, ed. Richard Bauckham (Edinburgh: T&T Clark, 1999), 230-31. For Bauckham’s sharp criticism, see 179-80. 55 Moltmann, Sun of Righteousness, 196. 53

164

Stephen N. Williams

a­ dvanced our knowledge of the world and that there are scientific truths that theology always has to take into account and, where relevant, explicitly incorporate into its work. Doctrines of creation must therefore assume, for example, the Big Bang and the broad evolutionary account of the world. In light of these statements, we should note that anything that is truly known by us in any sphere of life logically constrains any claims that we can validly make about anything else.56 How theologians relate theology and science is not always easy to say, because they may be related in tacit ways as, for example, in the case where the most die-hard defenders of the preeminence of biblical or theological truths will automatically adopt a heliocentric view of the world on nontheological grounds. To all appearances, Moltmann himself does not let science get in the way of conclusions reached by theology operating in its own right with what he judges to be its own criteria. We may generalize in relation to Moltmann’s thought what Richard Bauckham said about his treatment of creation and evolution: “The theological structure . . . is not without input from the natural sciences, but is not determined by them.”57 Of course, scientific influence on theology can be subtle even when disavowed. However, the direction of travel in Moltmann’s thought is set by theological considerations, and science exercises minimal constraints on those, playing, where appropriate, a supporting role.58 Until we arrive at Science and Wisdom, it is difficult to gauge the usefulness of trying to read Moltmann’s work according to the principle that revelation must open up new horizons for scientific interpretation. When Moltmann makes this statement, he is obviously not thinking of anything within the absolute sphere of revelation but only that which, in principle, touches the sphere of science. Science and Wisdom does seem to supply some warrant for the claim that this describes how Moltmann proceeds some of the time, but even here it is not fruitful as a comprehensive principle for interpreting his work. I think we are on better If I make a knowledge claim that turns out to be mistaken, then it was not known in the first place; I only thought that I knew it. “X is true” is analytic in the claim “I know x.” 57 Richard Bauckham, The Theology of Jürgen Moltmann (Edinburgh: T&T Clark, 1995), 190. 58 For the claim that Moltmann advances a “theological dictatorship over science,” see Buxton’s reference to Jim McPherson’s criticism in “Moltmann on Creation,” 43n11. 56

Jürgen Moltmann (1926–)

165

ground for understanding how Moltmann relates theology and science with reference to creation when we look at the particular topics that he handles rather than broad statements of principle. An example of this comes in his essay on the hermeneutics of nature in Sun of Righteousness. He also gnomically says elsewhere, “In the sphere of hermeneutics, scientists and theologians meet”—both groups are interpreters of nature.59 Scientific knowledge of nature is inadequate for properly understanding nature; understanding of the whole is needed if that knowledge is to be fruitful. “The ‘book of nature’ can only be read if we regard nature not as a world of facts but as a world of meanings,”60 and theology supplies the ultimate meanings. Theology has a long history, going back to the Syrian church fathers and the Cappadocians, of thinking in terms of two books—Scripture and nature.61 The book of nature is a book of signs; nature is sacramental. And, of course, the great sacrament manifest in the Christian church is the Eucharist. By virtue of the presence of the creative, life-giving Spirit, the presence of God will be perceived in all things, just as the body and blood of Christ is present in bread and wine, and the whole Christ is present in the whole celebration of the Supper. These are the traces of God, vestigia Dei, in the history of nature and civilization. These are the correspondences between created beings and their creator, the anticipations of their future true form or gestalt.62

The sources for our understanding of creation, then, are Scripture, theologically appropriated, and nature, scientifically described. Science can show us when scriptural tenets are outmoded. But it is theology that interprets the deep meaning of nature. Moltmann is more interested in this than in detailed theological engagement with science. On this score, it is instructive to notice what Moltmann does in his ongoing references to the German neo-Marxist philosopher Ernst Bloch. Bloch arrived on the scene early in Theology of Hope, and almost forty years later Moltmann could call Bloch’s Principle of Hope, the work that had engaged him then, Moltmann, “Is the World Unfinished?,” 404. Moltmann, Sun of Righteousness, 198. 61 For natural theology, see God in Creation, chap. 3. 62 Moltmann, Sun of Righteousness, 202. For elaboration of the claim that “it is only in the light of the Lord’s supper that the interpretation of ‘the signs of the times’ is possible,” see Jürgen Moltmann, The Church in the Power of the Spirit (London: SCM Press, 1977), 242-60, quotation from 243.

59 60

166

Stephen N. Williams

“still amazing” and aver, “From him I learnt to seek the divine mystery not just up above us in heaven, and not just within us in the ground of our being, but before us, and ahead of us, on our way into the unknown future.”63 In God in Creation, Bloch appears where we might expect references to scientists to appear—namely, at a point where Moltmann contrasts the mechanistic model of the world as a machine with the world as an open system.64 The same pattern of expectation and (relative) nonfulfillment appears later. Thus, when we read the first of Moltmann’s two essays on time in Science and Wisdom, we anticipate scientific discussion when we encounter an early remark on relating the time measurements of physics to subjective experiences of eternity. Yet it is to Bloch rather than the physicists that Moltmann turns. In a brief section on “Modes of Time and Modalities of Being,” Moltmann proposes to follow Bloch in relating “modes of time and modalities of being to each other.”65 Moltmann does not engage science except in a general way; Moltmann and Bloch are both occupied with the relationship of potentiality to actuality from a broadly philosophical point of view.66 Again, in his discussion “The Times of History” in Christian Eschatology, no sooner has Moltmann introduced the second law of thermodynamics and irreversible time than Bloch is invoked.67 In his later essay “Natural Science and the Hermeneutics of Nature,” Bloch soon appears when the question is posed of whether “the natural sciences do violence to nature,”68 so it is not the shadow of a scientist that is cast over the discussion. For evidence that Moltmann draws comparatively little on scientists at the very point Moltmann, Sun of Righteousness, 101. Moltmann, God in Creation, 315. Bloch appears elsewhere in God in Creation. 65 Moltmann, Science and Wisdom, 89. 66 Bloch could be taken as merely one example of the point that I am making, for Moltmann also refers to another philosopher or theologian at this point, Georg Picht, who also features in Moltmann’s work from Theology of Hope to Science and Wisdom. But Bloch’s profile is still higher. See, e.g., Jürgen Moltmann, “Ernst Bloch and Hope Without Faith,” in The Experiment Hope (London: SCM Press, 1975), 30-43; and Jürgen Moltmann, “Messianic Atheism,” in On Human Dignity: Political Theology and Ethics (London: SCM Press, 1984), 173-88. See also the appearance of Bloch (and Picht) in Moltmann’s discussion of “The Times of History” in Christian Eschatology, 284-92. 67 See Moltmann, Christian Eschatology, 284-92. Picht features here too (286). Bloch reappears in this section (290). 68 Moltmann, Sun of Righteousness, 191. 63 64

Jürgen Moltmann (1926–)

167

where we should expect to meet them, we could do worse than track his use of Ernst Bloch. Conclusion Although I have discussed some expectations arising from the surface of Moltmann’s literature, it seems to me that in at least three general respects we should not quarrel with the way in which Moltmann understands the principled relationship between theology and science in his pursuit of the doctrine of creation. First, he sustains a vision of the unity between science and theology forged by wisdom; second, he believes that it is the office of theology to provide a constructive critique of scientific culture; and third, he insists on locating the scientific enterprise in the one history of the one world. While those acquainted with the work of T. F. Torrance on theology and science will appreciate the difference between the way that Moltmann and Torrance approach their engagement, we should note Moltmann’s appreciation of his colleague’s resolute insistence on grappling with the relationship between theology and science.69 In sum, keeping those three points in mind, Moltmann’s distinctive contribution to the doctrine of creation at the intersection of theology and science lies not in any statements about the principles of interaction between the two disciplines but in his particular theology of zimzum. While he does explore questions in relation to time as well as to space, I have given examples where he leaves these explorations largely undeveloped at the expense of securing the wider picture about the history of nature. In my judgment, it is impossible to be dogmatic on whether this is a weakness. It is questionable whether we should possess a standard set of expectations of a theologian in the way of scientific engagement when he or she comes to explore creation. As long as a theologian does not proceed in flagrant contravention of scientific knowledge or ignore prevalent scientific theory, we must surely be flexible in our expectations with regard to the extent and nature of the theologian’s grasp of science. Theology has its own things to say just as science has its own things to say. Admittedly, the scientist-theologian or theologian-scientist is clearly at an advantage, but See Moltmann, Science and Wisdom, 204n23; and Moltmann, God in Creation, 347n27.

69

168

Stephen N. Williams

those of their number are not the only fruitful contributors to an understanding of creation. Moltmann is not of their number, but he is able to make telling connections between theology and science when and where he needs to. Further, Moltmann cannot be accused of lacking theological ambition in relation to science. His chapter “The Space of Creation” in God in Creation includes a brief discussion of the problem of absolute space, where he opines that Leibniz and Newton’s disagreement over this issue requires a theological intervention to settle it. We need to “think of creation as the mediation between the relative space of objects and the eternal space of God. It is only the concept of creation which distinguishes the space of God from the space of the created world.”70 Critical questions in respect of Moltmann’s understanding of creation arise less from his engagement with science than from his views on God, space, and zimzum, the area of his most detailed contribution to a theology of creation. There is a twofold vulnerability here: Moltmann is theologically vulnerable to both biblical and philosophical critique of what he is doing and scientifically vulnerable to any concepts of space that scientists may develop in the present or future which are both intrinsically plausible and impinge on Moltmann’s speculations. Yet Moltmann’s observations on the role of theology in relation to the Leibniz-Newton disagreement suggests that his theological confidence extends beyond the narrower sphere of engagement with the doctrinal tradition to the wider sphere of scientific adjudication. Is such confidence warranted? In principle, perhaps, as regards the potential scope of theological competence. Whether it is so in practice, as in Moltmann’s case, depends on an examination of God, divine space, and zimzum that I have not the human space to offer.

Moltmann, God in Creation, 156.

70

Jürgen Moltmann (1926–)

169

Recommended Reading Primary Works Moltmann, Jürgen. God in Creation: An Ecological Doctrine of Creation. London: SCM Press, 1985. ———. Science and Wisdom. London: SCM Press, 2003. ———. The Trinity and the Kingdom of God: The Doctrine of God. London: SCM Press, 1981.

Secondary Works Bauckham, Richard. Moltmann: Messianic Theology in the Making. Basingstoke, UK: Marshall Pickering, 1987. ———. The Theology of Jürgen Moltmann. Edinburgh: T&T Clark, 1995. Chung, Sung Wook, ed. Jürgen Moltmann and Evangelical Theology: A Critical Engagement. Eugene, OR: Pickwick, 2012. Torrance, Alan J. “Creatio Ex Nihilo and the Spatio-Temporal Dimensions, with Specific Reference to Jürgen Moltmann and D. C. Williams.” In The Doctrine of Creation: Essays in Dogmatics, History and Philosophy, edited by Colin E. Gunton, 83-103. New York: T&T Clark, 1997.

8 WO L F HA RT PA N N E N B E R G ( 192 8 – 2 0 14 ) Nature, Contingency, and the Spirit C h ristoph S ch wöb el Wolfhart Pannenberg, baptized as an infant in the Lutheran Church, grew up in an agnostic family in Stettin (now Szczecin), Poland. After an intense religious experience as a teenager, Pannenberg read widely in philosophy and theology in order to refute the atheistic worldview he had encountered in paradigmatic form in the philosophy of Friedrich Nietzsche. He found that Christian faith offered the best orientation for life. After being drafted into the army at the age of sixteen and a short period as a British prisoner of war, he became a theologian to find reasons for this conviction. Pannenberg studied theology and philosophy at the universities of Berlin, Göttingen, Heidelberg, and Basel, where he studied for a semester with Karl Barth. From the publication of the programmatic collection of essays Offenbarung als Geschichte in 1961 (Revelation as History, 1968), Pannenberg’s theology was perceived as an alternative to the prevailing Word of God theologies of Karl Barth and Rudolf Bultmann. From the early 1960s Pannenberg was invited frequently to the United States, and his theology was soon more discussed in the English-speaking world than in his native Germany. Together with Jürgen Moltmann, Pannenberg became one of the first European theologians of the late twentieth and early twentyfirst century to be globally discussed, and he was the recipient of numerous honorary doctorates.

T

he engagement with the natural sciences, their findings, their theories and theory formation, and their effect on our understanding of reality was an ongoing concern of Wolfhart Pannenberg’s theology throughout his career. He himself dated the beginning of this engagement back to the 1950s and the conversations between natural scientists and theologians in Göttingen, established in 1949 by the physicists Günter

Wolfhart Pannenberg (1928–2014)

171

Howe and Carl Friedrich von Weizsäcker.1 These conversations continued in Heidelberg in a working group at the research academy of the Protestant Church in Germany at Heidelberg. In this context, the paper “Contingency and Natural Law” (1966) was written, which shaped much of Pannenberg’s later reflections on a theology of nature.2 Pannenberg pursued these questions first in conversation with theoretical physics; later his engagement with the sciences expanded to the philosophy of science and, especially in the context of his Anthropology in Theological Perspective, to anthropology and human biology. The theme of contingency, crucial for the understanding of God’s action in history and central for Pannenberg’s programmatic proposal presented in the collection of essays by the “Pannenberg Circle” with the programmatic title Revelation as History,3 shows that the central questions of Pannenberg’s engagement with the natural sciences are also those that are central to the development of his own theology. One can interpret the logic of the development of Pannenberg’s theology as filling in the “missing links” of his conception of revelation as history.4 Pannenberg’s engagement with the natural sciences parallels this continuous extension of the scope of his theology until he presented it in its finished form in his Systematic Theology. Pannenberg’s theology is therefore one of the few examples of a theology, perhaps the only one in the twentieth century, in which all crucial steps toward the formation of his “system” were accompanied by a parallel engagement with the natural sciences, their philosophical implications, and the questions they raise for philosophy and theology. From the beginning, Pannenberg aims at conducting a theological conversation with the sciences, an endeavor quite different from the science and religion dialogues, as they are conventionally perceived. He is less concerned with offering an apology for religion that might be Cf. Günter Howe, Gespräch zwischen Theologie und Physik, Glaube und Forschung. Veröffentlichungen des Christophorus-Stiftes in Hemer, Bd. 2 (Gladbeck: Freizeiten Verlag, 1950). 2 In Wolfhart Pannenberg, Towards a Theology of Nature: Essays on Science and Faith, ed. Ted Peters (Louisville, KY: Westminster John Knox Press, 1993), 72-122. 3 Wolfhart Pannenberg, ed., Revelation as History (London: Sheed and Ward, 1969). 4 On this, see Christoph Schwöbel, “Wolfhart Pannenberg,” in The Modern Theologians: An Introduction to Christian Theology in the Twentieth Century, 2nd ed., ed. David F. Ford (Oxford: Blackwell, 1997), 180-208. 1

172

Christoph Schwöbel

acceptable to scientists. He inquires in which way the natural sciences call for clarifications in the theological understanding of God and his action in the world and in which way theological reflection offers suggestions for the interpretations of the findings and theories of the natural sciences. Furthermore, for Pannenberg philosophy is always the bridging discipline, one could almost say the simultaneous interpreter, translating in both directions, in the conversations of theology with the sciences. This implies, on the one hand, that he attempts to make explicit the philo­ sophical presuppositions and implications of scientific theories that in the sciences often remain implicit. On the other hand, he is also concerned to point to the philosophical implications of Christian beliefs, including the ways in which Christian faith calls for conceptual revisions of philosophical conceptions and redescriptions of scientific theories. The conversations between theology and the sciences, which are often conducted through philosophy as the interpreter, are in Pannenberg’s work always traced back through their histories. The history of conceptformation in theology, philosophy, and science is the medium for Pannenberg’s explorations.5 This often challenges received views of all three histories involved and points to new combinations between them, new possibilities opened up by the new interpretations that Pannenberg offers. In these attempts, recourse to the biblical witnesses is always the starting point for the analysis of the concept formation of Christian theology, often calling for revisions of received conceptualities, as well as a reference point for exploring the adequacy of philosophical concepts and scientific theories. In the following reflections I shall briefly consider Pannenberg’s understanding of nature, his understanding of contingency and its complex relations to necessity and possibility, and, finally, his most innovative interpretation of the being and operation of the Spirit as a field of force. Pannenberg’s achievement in all three areas of investigation raises a few questions that seem to me profitable to pursue in conversation with his thought. 5

This applies to all aspects of Pannenberg’s thought. The histories of theological or scientific concepts and theories are an indispensable element of conceptual analysis, and they point forward to their resolution in the future.

Wolfhart Pannenberg (1928–2014)

173

Understanding Nature From a theological perspective, it seems surprising that Pannenberg starts his engagement with the relationship between contingency and natural law with considerations on the possibility of a theology of nature. However, this starting point is programmatic. Pannenberg observes that in modernity a rift has opened between theology and the knowledge of nature. For Christian faith, this chasm seems unacceptable if the God of Christian faith is also the Lord of nature. Pannenberg fears that in this context, marked by the development of the modern sciences, starting theological reflections from the concept of creation already accepts the separation of theology and the sciences by beginning with a purely theological concept. Of course, he does not want to give up the understanding of creation. However, rather than serving as a starting point it marks for him the possible result of a theology of nature, of a theological interpretation of natural reality. Furthermore, Pannenberg fears that the term creation might focus the discussion exclusively on the beginning of the cosmos. In contrast to that, a theology of nature would have to look at nature in the totality of its processes and try to relate them to the reality of God. Pannenberg does not deny that the development of the modern sciences is in large part the history of the emancipation of the natural sciences from their theological presuppositions.6 He concedes the defeat of theological apologetics that had tried to place the action of God only in the gaps of the knowledge of nature in the sciences. It is, in Pannenberg’s view, unsatisfactory to respond to the defeat of theological apologetics with a cautious attitude that attempts to develop a theology of creation exclusively on the theological plane that remains unassailable by the criticism of the sciences—an approach that Pannenberg sees exemplified in Karl Barth’s doctrine of creation.7 Such an approach makes theology Pannenberg has traced this development numerous times. Perhaps the most comprehensive is the essay “God and Nature,” in Theology of Nature, 50-71. 7 Pannenberg frequently refers to Barth’s remark in the foreword of CD III/1: “There can be no scientific problems, questions, objections or aids in relation to what Holy Scripture and the Christian Church understand by the divine work of creation.” Karl Barth, Church Dogmatics, trans. G. T. Thompson et al., ed. G. W. Bromiley and T. F. Torrance (Edinburgh: T&T Clark, 1936–1977), III/1:ix. 6

174

Christoph Schwöbel

irrelevant not only for the sciences but also for all people whose understanding of reality is largely formed by the sciences. In contrast, Pannenberg states that nature as it is investigated by the natural sciences should be claimed by a theology of creation. He equally rejects the attempt to reduce the doctrine of creation to the view that “God has created me,” making the subjective experience of creatureliness the beginning and end of a theological engagement with nature.8 Pannenberg is clear that the idea of God can no longer be developed on the basis of the knowledge of nature. For him, anthropology is the ground on which the function of the idea of God for a human understanding of reality must be analyzed.9 This, in turn, might provide reasons for the justification of the idea of God. However, even if the idea of God cannot be developed on the basis of our knowledge of nature, it must nevertheless be shown to have validity for the whole of human experience of the self and the world. A God who is not the origin and consummator of the world of nature could not be the power that determines everything and so could not be God. Pannenberg’s constructive aim is a “common ground” to which theological and scientific problems can be related without obscuring the difference between both forms of thought.10 Constructively, Pannenberg develops his understanding of nature on the basis of what he calls the Israelite understanding of God who acts in events that happen contingently. On this view, only the future will show the full meaning of everything that happens contingently. Can this understanding of the contingency of events be constructively applied to the understanding of nature in the natural sciences? Pannenberg especially has the discipline of physics in mind. Prima facie the very concept of a law of nature seems to contradict this possibility since it seems to establish a necessary connection between events A and B that can be formulated as a general regularity, a universal hypothesis, we might say, and For Pannenberg, this existential approach to the doctrine of creation is exemplified by Rudolf Bultmann, Friedrich Gogarten, and their schools. 9 This is a continuing characteristic of Pannenberg’s approach from the early What Is Man?, trans. Douane A. Priebe (Philadelphia: Fortress Press, 1971) to Anthropology in Theological Perspective, trans. Matthew J. O’Connell (Edinburgh: T&T Clark, 1985). 10 Pannenberg, Theology of Nature, 76. 8

Wolfhart Pannenberg (1928–2014)

175

so supports the deterministic view of nature in classical physics.11 Pannenberg clearly sees that this kind of determinism would establish a realm of nature that would, in principle, not be accessible to God’s action. A deistic God who establishes the laws of nature at the beginning and then takes early retirement—what Leibniz called “the God of the sabbath”—seems to be the only way in which language of God can still function in this framework. Once inner-worldly necessity of natural laws is established, there is no relationship between God and nature. Pannenberg attempts to create space for a view of God as acting in and interacting with the world of nature in its particular events and in its general regularities. Therefore, he attempts to show that the contingency of individual events, of their regularities and of the direction of the whole process of nature, is the framework within which the regularities of the connection between particular events as well as particular events themselves have to be understood.12 This view of contingency serves as the common ground to which scientific and theological statements must be related. The motivation of attempting to show that contingency is the framework for understanding nature can be justified by pointing to the indeterminacy of elementary events on the quantum level, instabilities that appear in chaotic processes, and the irreversibility of the course of time as it is stated in the second law of thermodynamics.13 Pannenberg For Pannenberg, this deterministic view of the laws of nature is not so much a result of classical physics as a “sort of religious faith, a late form of Greek cosmos piety,” even a “sort of illusion” that may have been revealed by the much more realistic views of modern physics and by the philosophical difficulties in giving a precise content to the term laws of nature. Today one would want to differentiate between deterministic relationships between cause and effect that can be expressed as mathematical functions as in the laws of mechanics and electrodynamics, statements about statistical means as we know it from thermodynamics, and statements referring to collective probabilities as in quantum mechanics or the deterministic-chaotic behavior that we find in emergent processes which are seen as self-organizing. This differentiation does not solve the problem of the philosophical interpretation of “laws of nature” whether they are descriptions of observed regularities of fundamentally arbitrary processes, so that the world is a mosaic of singular facts which may generate patterns that cannot be seen as necessary connections between facts, or whether they express the power to necessitate an effect from a given cause (the so-called universal theory). Pannenberg is clearly on the side of the regularities (or systems) theory espoused by David Lewis. For an exhaustive account of the philosophical debate, see John W. Carrol, “Laws of Nature,” in The Stanford Encyclopedia of Philosophy (Fall 2016 Edition), ed. Edward N. Zalta, https://plato.stanford.edu/archives/fall2016/entries/laws-of-nature/. 12 Cf. Pannenberg, Theology of Nature, 81-86. 13 Pannenberg time and again refers to Carl Friedrich von Weizsäcker, The History of Nature (Chicago: University of Chicago Press, 1949), as the key reference text for the cosmological and metaphysical implications of the second law of thermodynamics. 11

176

Christoph Schwöbel

does not question that the overwhelming majority of the sequences in the processes of nature are characterized by uniformities that can be stated in the universal hypotheses of the sciences. However, Pannenberg insists, these uniformities occur with regard to contingent sequences of events of which every single one is contingent so that each event is followed by a contingently new event. Because of the irreversibility of time, every event is a singularity and is contingent so that their sequence has the form of a history. Contingency refers to that which factually exists but exists neither necessarily nor necessarily not. In other words, everything that is not impossible (i.e., that is not necessarily not) and exists, although it could also not exist, is contingent. The realm of contingency comprises both that which is strictly singular and that which is uniformly governed by the “necessity” of natural laws. “Contingency means here . . . always: that which is not necessary on the basis of what is past,” says Pannenberg, and he dubs this understanding of contingency “historical contingency.”14 Once this understanding of contingency is established, it becomes clear that contingency is the category comprising both nature and history. In nature it is the foundational term for singularities as well as for rule-governed uniformities. This understanding of nature is open for God’s relationship to the world both in the uniform regularities of nature and in the contingent events in nature. God’s creative action maintains the regularities of the natural laws in order to provide the basis for the emergence of independent creatures. On this view, natural laws are strictly relative to God’s creative agency; they are, one could say, consuetudines Dei, habits of God the Creator, as Leibniz has remarked.15 Once contingency is established as the framework for understanding both singularities and uniformities, Pannenberg can proceed with developing his understanding of nature as history, in conversation with Carl Friedrich von Weizsäcker’s classic Die Geschichte der Natur (1948).16 The Pannenberg, Theology of Nature, 116n11. On Leibniz’s theology, see Wenchao Li and Hartmut Rudolph, eds., Leibniz (Stuttgart: Franz Steiner Verlag, 2013). 16 To what extent Pannenberg’s theology of nature is an ongoing conversation can be seen from the essays collected in The Historicity of Nature: Essays on Science and Theology, ed. Niels Henrik Gregersen (West Conshohoken, PA: Templeton Foundation Press, 2008). 14 15

Wolfhart Pannenberg (1928–2014)

177

notion of contingency has far-reaching implications for the understanding of the historical process of nature—both for theology and for natural science. For Pannenberg, the assumption of a goal-directedness from the beginning that governs every event and so establishes continuity is incompatible with this notion of contingency since it would imply that every new event can be understood from its antecedent conditions. If there is to be continuity in contingency, it must be understood as a kind of backward continuity, a retroactive continuity in which what happens later establishes the continuity with what happens earlier. If this con­ tinuity is to establish not only the continuity in different strands of events but is also to be seen as constitutive for the unity of everything that occurs in nature and history, it must be seen as the consummation of everything that occurs in God’s perfecting agency. For Pannenberg, this is a distinctive Christian thought since it is the basic conviction of Christian faith that the end of history has already in advance—proleptically is the term he uses—occurred in Jesus Christ. Pannenberg is doubtful whether the idea of the end as the completion of history, which is the basis for conceiving the unity of all processes in nature and history, can be abstracted from its theological roots. This idea, however, cannot be stated by theology in terms of what one could call a pseudophysical rival theory in the sciences; it must be stated as a theological view, based on theological considerations. In the conversations with the natural sciences it can be used only as a heuristic tool for exploring the question of the unity of all events and processes. A theology of nature is thus the theological presupposition of conceiving the unity of nature. Reclaiming Contingency: Lessons from Conceptual History Pannenberg sees contingency as the basic category for developing an understanding of nature that can serve as the mutual reference point for both the natural sciences and Christian theology.17 To strengthen this claim, he offers a complex genealogy for the loss of significance of the category of This is a concern Pannenberg shares with T. F. Torrance. Cf. Torrance’s works Space, Time and Incarnation (London: Oxford University Press, 1969) and Divine and Contingent Order (Oxford: Oxford University Press, 1981).

17

178

Christoph Schwöbel

contingency and of the possibilities of its retrieval, already contained in the history of Christian thought. The Aristotelian heritage of Christian theology had the effect, Pannenberg claims, that contingency was predominantly associated with something material that is actual but could possibly also be different. Contingent is that which does not essentially belong to the concept of something but can be assumed or received by it. The association of contingency with actual matter that could be different is expanded in the Middle Ages (e.g., in Thomas Aquinas) to the contingent result of an act of choosing. Pannenberg ascribes a reevaluation of the concept of contingency to his philosophical and theological hero in medieval times, John Duns Scotus. Only a freely acting cause can effect something contingent, because everything that effects something out of the necessity of its nature can, logically, not cause something that could also be something different. The first cause of everything effects something contingent not out of the necessity of its nature but by the freedom of its will. However, if God is immutable, how can he effect something contingent? While Thomas Aquinas had already located the contingency of creation in the will of God, Duns Scotus radicalizes this view by emphasizing the synchronic contingency of everything as rooted in God’s freedom. According to Duns Scotus, a created will can choose contradictory things only in successive acts. That is due to the imperfection of created agency. The eternal will of God, however, comprises all contents of his choosing in one eternal immutable act, even if these aims of God’s willing are actualized in different times and places in the created world. Pannenberg observes that this reevaluation of contingency is due to a different understanding of possibility. The possibility of whatever is precedes its actualization in the divine intellect. This possibility is the presupposition of God’s choice, and whatever God chooses to actualize is contingently actual. Contingency is no longer located in the indeterminacy of matter (as in Aristotle and Aquinas) but in the freedom of God, whose free will is the ground of everything that exists contingently. The relationship established between God’s freedom and contingency, following Pannenberg’s interpretation of Scotus, enables a paradigm shift.18 For a theological discussion of the shift from the abstract to the particular whose contingent particular existence and characteristics are rooted in the divine will, see Colin E. Gunton, The

18

Wolfhart Pannenberg (1928–2014)

179

The task of reclaiming contingency also involves understanding the abolition of contingency in the mechanistic theories of the sciences in the eighteenth and nineteenth century and analyzing the philosophical and theological presuppositions of such views. In the way Pannenberg variously tells the story, Descartes is, for the best of theological intentions, the main culprit. Descartes’s book Le Monde—written around 1630 but only published in 1664, after his death, because he feared the Inquisition in the nervous atmosphere after the second trial of Galileo—is Pannenberg’s main piece of evidence for this view.19 In Pannenberg’s interpretation, which rests on Alexandre Koyré’s Newtonian Studies,20 the crucial step in Descartes’s work is the formulation of the principle of inertia. According to this principle, every body, every part of matter stays in the state in which it is unless it is moved from without. On the basis of this principle all changes in the state of material bodies in the world can be ascribed to the mutual interaction of material bodies. Recourse to God is no longer necessary for interpreting particular events, only for creating bodies, the principle of inertia, and conserving them in existence. Since in Descartes’s view bodies are in motion from the beginning and have an intrinsic tendency to continue this movement on a straight trajectory, any deviation from this straight path is caused by the interaction of bodies with one another. The nonintervention, even the noninvolvement of God implied in this view can be justified theologically by the immutability of God. The principle of inertia, applied to moving bodies, in this way also destroys, as Pannenberg acutely observes, the foundation of Aquinas’s proof for the existence of God in the first of his five ways (ex parte motu).21 In Pannenberg’s interpretation, Spinoza completes the abolition of contingency by a deterministic network of natural causes that is established in its necessary order in God’s action and is not open to divine intervention or interaction. In Spinoza, Pannenberg maintains, the view that motion is proper to the bodies themselves and Triune Creator: A Historical and Systematic Study, Edinburgh Studies in Constructive Theology (Edinburgh: Edinburgh University Press, 1998), 117-21. For an exposition emphasizing the position of John Duns Scotus in medieval philosophy, see Kurt Flasch, Das philosophische Denken im Mittelater: Von Augustin zu Macchiavelli (Stuttgart: Philipp Reclam, 2000), 482-98. 19 Cf. Pannenberg, Theology of Nature, 53-55, 60-64. 20 Alexandre Koyré, Newtonian Studies (Cambridge, MA: Harvard University Press, 1965), 66-73. 21 Cf. Pannenberg, Theology of Nature, 54.

180

Christoph Schwöbel

does not require God’s interaction and the view that, in any case, the immutability of God prevents such an intervention form the foundations of a mechanistic view of reality. In these genealogical explorations, Pannenberg shows that science often rests on tacit philosophical assumptions and is bound to proceed as if these are not philosophical assumptions but scientific data. For him, the best way to make the data again accessible to interpretation is to dissociate them from the philosophical views in which they seem to be quasi-naturally embedded. For the philosophical retrieval of the notion of contingency, the controversy between Leibniz and Newton and Newton’s apologist Samuel Clarke is of special significance.22 Pannenberg rejects Leibniz’s criticism that Newton’s characterization of space as the sensorium Dei, the sensory apparatus of God, turns God into a material being. Instead, following the interpretation of Newton’s friend Samuel Clarke, he understands the sensorium not as an organ of perception but as the medium of the generation of all things. Just like our sensory apparatus produces the images of things, so God produces through the mediation of space the things themselves.23 In Pannenberg’s interpretation, limitless space must in Newton’s view both be seen as an implication of God’s immensity and as an effect of God because it offers space for the creatures given its divisibility. The undivided and infinite space, God’s immensity, is thus in Pannenberg’s interpretation of Newton the presupposition of all divided geometrical spaces, and as such it is the medium of God’s presence with his creatures without compromising God’s transcendence. Pannenberg has a similar argument, this time developed on the basis of Plotinus’s reflections on eternity, for God’s eternity as the undivided totality of time functioning as the presupposition of all notions of periods of time and their succession.24 Space and time are—in Pannenberg’s interpretation of Newton’s (and Plotinus’s) view—modes of God’s presence in creation. Pannenberg, Theology of Nature, 59-63. Pannenberg’s account is here also dependent on Alexandre Koyré, From the Closed World to the Infinite Universe (Baltimore, MD: Johns Hopkins University Press, 1956). 24 The most detailed form of the argument can be found in Wolfhart Pannenberg, Systematic Theology, vols. 1–3, trans. Geoffrey W. Bromiley (Grand Rapids, MI: Eerdmans, 1992–1998), 3:401-10, esp. 403, with reference to Plotinus, Enn. 3.7.11 and repeated in the eschatology of vol. 3. 22 23

Wolfhart Pannenberg (1928–2014)

181

With this interpretation, which is not only a matter of historical reconstruction but of using the historical debates as a means for the establishment of systematic possibilities, Pannenberg has secured a metaphysical interpretation of Newton’s concept of space that he claims remains valid despite the criticism of modern physics. That is, the corrections in modern physics to Newton’s view of space do not effect the philosophical content. Similarly, he claims that Plotinus’s understanding of eternity as the ground of the possibility of all structured times, the understanding of the duration of time and of location in time, still is the most elaborate understanding of the relation of eternity and time in Western philosophy. Both concepts of time and space find their proper relationship in the interpretation of the Spirit as the field of force through which God is present to his creatures. The Spirit as a Field of Force Pannenberg’s interpretation of the Spirit, for which he always refers to John 4:24 (“God is spirit, and his worshipers must worship in the Spirit and in truth”), is without doubt the most innovative but also the most daring proposal he has introduced into theology, not only into the theology of nature but most of all into the doctrine of God. It is also the one that Pannenberg employs in order to tackle what he sees as the most serious problem of the mechanistic interpretation of the processes of nature. If all effects of force are caused by material bodies and their interaction, there is no room for God in the processes of nature. God is politely asked to leave the world of nature. For Pannenberg, a constructive way out of this situation is to appropriate the understanding of the total field of natural occurrences as a force field (appearing in the material shapes of its singularities, as it was developed by Michael Faraday) and to apply it to the understanding of the Spirit in the biblical traditions. For Faraday (1791–1867), the attempt at finding in his time a unified theory of natural forces that could be comprehended through his idea of lines of force, constituting a force field, were not independent of his religious beliefs as a follower of the Sandemanian strand of the Scottish awakening.25 With the idea of a force Cf. Geoffrey Cantor, Michael Faraday, Sandemanian and Scientist: A Study of Science and Religion in the Nineteenth Century (London: Macmillan, 1991).

25

182

Christoph Schwöbel

field, Faraday reversed the order of the relationship of body (mass) and force. Force is not exercised between already existing bodies. Rather, body and mass are “secondary phenomena a concentration of force at particular places and points of the field.”26 This notion is attractive for Pannenberg not only because it transcends the expulsion of God from a mechanistic universe of the interaction of mass but also, in keeping with Faraday’s vision of a single comprehensive force field, because it claims the priority of the whole over its parts.27 Theologically, Pannenberg deviates from the interpretation of John 4:24 in the theological tradition since Origen, who interpreted the Spirit in the sense of the nous, the immaterial rational principle. Against such a view, Pannenberg points to the connection between field of force and the Stoic notion of pneuma, spirit, that Max Jammer had hinted at. The Stoic notion of the pneuma is the all-pervasive air that by its one tension (the Greek word is tonos, as in muscle tone) holds the whole cosmos together, which is the cause of all motions and of all specific qualities of substances.28 This is for Pannenberg very close to the Hebrew understanding of ruach, variously translated “air,” “breath,” and “wind.” There is for him also a good trinitarian reason for keeping close to that Hebrew understanding if we understand the divine ousia as the field that is mani­ fested in the “persons” of Father, Son, and Spirit. The person of Spirit is not himself this field but has to be understood as a singularity of the field of the divine ousia. However, there is a close connection between the role of the Spirit as the mediator of communion within the Trinity and as the mediator of relationship between the Trinity and the created world. Applied to created reality, force no longer has to be understood as an effect of a body and its mass. Conversely, if we follow the intuitions Pannenberg, Theology of Nature, 35. Pannenberg refers to William Berkson, Fields of Force: The Development of a World View from Faraday to Einstein (New York: John Wiley & Sons, 1974). For a concise overview of the issues involved in the notion of a field of force, see Mary B. Hesse, Forces and Fields: The Concept of Action at a Distance in the History of Physics, Philosophical Library (Edinburgh: Thomas Nelson and Sons, 2008). For the importance of the relationship of part and whole, see Pannenberg’s paper “Theology and the Categories ‘Part’ and ‘Whole,’” in Metaphysics and the Idea of God, trans. Philip Clayton (Edinburgh: T&T Clark, 1988), 130-52. 28 Cf. Max Jammer, “Feld, Feldtheorie,” in Historisches Wörterbuch der Philosophie, ed. Joachim Ritter (Darmstadt: WBG Academic, 2019), 2:923-26. Jammer sees the Stoic concept of pneuma as a “direct predecessor” (p. 923) of the concept of a field. 26 27

Wolfhart Pannenberg (1928–2014)

183

of field theory, bodies are forms of appearance, manifestations of the field of force. If God is Spirit, and if the Spirit is to be understood as a field of force, how does that reintroduce God into the understanding of natural processes from which he had been evicted in the eighteenth century? Pannenberg explains this in exactly the same way as he had interpreted the relationship of God to space and time. Just as God’s undivided immensity is the condition for the existence of created spaces and the way in which God, in virtue of this constitutive relationship, is present in every created space and is also their interrelationship, and just as God’s eternity as the undivided totality of all time is the ground of the possibility of all times and of the connection between different points in time and periods of time, so God’s omnipotence as the force field of the Spirit is the condition for any created transmission of force and the “medium” within which they occur. From this concept of the Spirit as the field of force that is the ground of the possibility of all created action and reaction and the condition for the interaction of God with and in the created agencies, Pannenberg reinterprets the relationship of God to space and time. First, he reduces the concept of space to time as the simultaneity of different entities, events, and processes. Second, he interprets the relationship of God’s eternity to time as one where the future is understood both as the horizon of the contingency of every new event and the sphere of the unity of all that remains unfinished and incomplete in time. The operation of the Spirit in creation as the anticipation of the future consummation of everything (Rom 8:23) is therefore the precursor of the eschatological reality of everything. At first sight, there is a tension between theology’s view of God’s future as the origin of every event and the source of its possible perfection and the way in which the majority of the natural sciences depict all processes as beginning in the past and passing into the future. For Pannenberg, however, views focusing on the way the past shapes the present and views where the whole process is understood not to be “pushed” from the past but to be “drawn” from the future do not have to be incompatible. In the realization of events, which are originated from God’s future, there is the gradual building up of persistent structures

184

Christoph Schwöbel

that manifest God’s faithfulness to his creation. From the future, the constancy of God’s ways with his creation is continuously built up as the connectedness of events that make the knowledge of the sciences in lawlike regularities possible. In this way, the dynamic of the divine Spirit must be thought of in connection with time and space. It is in the field of force of the Spirit that the power of the Spirit grants the creatures their own presence and duration in time and consolidates their simultaneity in space. From the perspective of the creature, its advent, its origin from the future of the Spirit, presents itself as the descent from the past. The Spirit, however, encounters every creature as its future, which includes and integrates its past and present. In Conversation with Wolfhart Pannenberg Pannenberg’s theology of nature, which is part and parcel with his whole theological system both in its diachronic development and in its synchronic systematic presentation, is an impressive achievement in contemporary theology. It presents the most sustained attempt by any systematic theologian of the twentieth and twenty-first century so far to integrate the natural sciences in his overall theological enterprise and to show the creativity with which theology can relate—often through the uncovering of the basic concepts and root metaphors of science—to theory formation in the sciences. Its ethos of maintaining the unity of reality as God’s creation against the tendency to relegate theology to a special realm outside the interaction with the other spheres of human knowledge, and its insistence on the unity of truth in the face of all temptations either to suspend the question of truth altogether or to separate the truth of faith from what we hold to be true in all other spheres of life, makes Pannenberg’s theology particularly impressive. Pannenberg’s aim of achieving a “consonance” between our attempts at clarifying the intellectus fidei and the intellectus scientiarum is without reservation to be applauded. Against this laudable background, I would like to raise three questions. In search of a new paradigm. Pannenberg formulates his theological reflections within the paradigm of classical physics. Within this p ­ aradigm, we can describe the interaction of deaf and mute entities: force, mass, and impulse. For historical reasons, Pannenberg goes back to the point

Wolfhart Pannenberg (1928–2014)

185

where Descartes’s law of inertia forms the basis for the eviction of God and God’s action from the processes of nature. In this way, Pannenberg challenges the philosophical presuppositions of classical mechanics, both by his interpretation of contingency and his redefinition of the concepts of space and time, and he offers a novel interpretation for the being and interpretation of the Spirit by developing a philosophical interpretation of Faraday’s and later Clerk Maxwell’s understanding of a field of force. While this way of proceeding, of going back to where the problem started, can be justified as an attempt to unravel the knot of presuppositions that lead to the exclusion of God from nature, one could also ask the more radical question: Do we have to go forward to another paradigm in search of a more appropriate conceptuality to reflect on God’s action in relation to nature? My suggestion is that the paradigm of communication offers a more promising way of reflecting on the relationship between theology and the natural sciences. In this way we might also open up a discussion beyond Pannenberg’s predominant focus on theology’s relationship to physics to other natural sciences. Pannenberg’s initial theological program is offered as a counter­ proposal to the theologies of the Word that dominated the theological scene in various forms in German-speaking theology for much of the twentieth century. In Pannenberg’s view, revelation occurs indirectly through God’s action in history, and “the Word relates to revelation as prophecy, as guidance and as report.” This contrast of understanding God’s relationship to creation in terms of historical agency instead of God’s communication by the divine word is maintained even in Pannenberg’s Systematic Theology. Is the contrast between God’s action in history and God’s word a valid contrast? If we follow Pannenberg’s own theological method by looking first at the biblical witnesses, we find that God’s acting and God’s speaking are in most biblical traditions not viewed as an alternative. God’s word is understood as God’s efficacious word, as God’s speech act, by means of which God creates being and meaning and God’s agency is understood as God’s communicative action, communicating communion by establishing a communicative relationship with God’s creation. It is because God’s relationship to the world in creation, reconciliation, and the consummation is always a

186

Christoph Schwöbel

communicative relationship that the whole of creation responds to God’s address by praising the Creator. When we go back in the history of theology and science, again following Pannenberg’s example of retrieving the often submerged history of concepts for describing our relationship to God and the world, we find that the world of nature is understood as a book that is to be read, for Christians by the guidance of Scripture. This is not, at least not primarily, a paradigm of causality but a hermeneutic paradigm. The world is understood as a semiotic system where we try to read in order to understand the address of God to God’s creatures. What is the difference of the hermeneutic paradigm of reading from the causal paradigm of searching for relationships between causes and effects? Again, it is the early modern period that provides the most interesting example. Causal relationships do not convey meaning. Once the fourfold scheme of causes is reduced to efficient causality, we have only the relationship between event A and its effect B, whatever they may mean. The outcome of this is most clearly seen in Kant’s theory of experience. Our senses present us with the puzzling pluriformity of the sense impressions, which remain meaningless unless perceptions and concepts are brought together in the synthetic acts of experience. The divorce of being and meaning in the paradigm of causality is perhaps the greatest obstacle for a theological interpretation of nature in conversation with the sciences. Being is devoid of meaning unless we invest it with meaning. Not so in the hermeneutic paradigm of reading in the book of nature. Whatever has being has meaning in virtue of an author who invests everything that he creates with meaning. We do not create meaning but discover the meaning the Creator invested in creation from the ­beginning by reading the book of nature. Is this just an obsolete way of regarding nature that led to such fanciful ways of discovering meaning as the Kabbalah and astrology? Is not the parting of astrology and ­astronomy, which for people like Melanchthon and Kepler still belonged together, a sure sign of the victory of the causal paradigm over the ­hermeneutic paradigm?29 For an acute analysis of the scientific inquiry as a religious practice in early modern times, see Anne-Charlott Trepp, Von der Glückseligkeit alles zu wissen: Die Erforschung der Natur als

29

Wolfhart Pannenberg (1928–2014)

187

If we look at the development of the natural sciences in the twentieth century, we can detect an increasing turn from the mechanistic paradigm to the paradigm of reading. The discovery that processes of encoding and decoding form part of basic structures of life (for instance, in the human immune system) and the deciphering of human DNA, which has developed into reading the encoding and decoding in proteins, has offered many examples of the fruitfulness of working with the old and new ­paradigm of communication. Could it not be that large areas that were formerly analyzed exclusively in terms of causal relations are also open to being investigated as processes of information processing so that meaning structures have causal effects, perhaps also on the micro level of quantum mechanics and the macro level of cosmic evolution?30 “How to do things with words?” seems to be the maxim that can go through many variations in this paradigm shift. “How to effect events and processes through information?” and “How to create being from meaning?” are just two that immediately come to mind. The paradigm of communication has a long history, beginning with the biblical witnesses and continuing for a long time before and after the mechanistic era, which now appears as something of an interlude.31 Especially Reformation theology offers many striking examples of this way of understanding reality. “Thus the sun, the moon, the heaven and the earth, Peter, Paul, I and you etc. we are all words of God (vocabula Dei),” Martin Luther says in his exposition of Genesis. And he expands on that in the religiöse Praxis in der Frühen Neuzeit (Frankfurt: Campus, 2009). With regard to Kepler, see my paper “Der dezentrierte Kosmos und der zentrierte Mensch: Bemerkungen zu den theologischen Implikationen weltbildlicher Veränderungen im Anschluss an die Theologie Johannes Keplers (1571–1630),” in Der entgrenzte Kosmos und der begrenzte Mensch: Beiträge zum Verständnis von Kosmologie und Anthropologie, ed. Bernd Janowski and Christoph Schwöbel (Göttingen: Vandenhoeck & Ruprecht, 2016), 98-123. 30 John Polkinghorne has developed the view that changes in chaotic systems can be achieved through “active information” as a mode of top-down causality that does not violate the principle of energy conservation, especially in his books Belief in God in an Age of Science (New Haven, CT: Yale University Press, 1998) and Faith, Science and Understanding (New Haven, CT: Yale University Press, 2000). On the development of Polkinghorne’s view, see Ignacio Silva, “John Polkinghorne on Divine Action: A Coherent Evolution,” Science and Christian Belief 24 (2012): 19-30. 31 The history of the understanding of nature as a book and its relationship to the book of Scripture, from its beginning in Augustine through its development by Bonaventure, Raymond Lull, Nicolaus of Kues, Galileo, and Kepler to J. G. Hamann is documented in H. M. Nobis, “Buch der Natur,” in Historisches Wörterbuch der Philosophie, ed. Joachim Ritter (Darmstadt: WBG Academic, 2019), 1:957-59.

188

Christoph Schwöbel

same lecture series: “Every bird and fish are nothing else than words of the divine grammar.”32 Would it be a promising course to follow Luther’s ideas of creation as a divine speech act that presupposes his understanding of the Trinity as eternal conversation in which the Father represents the grammar, the Son the dialectics, and the Spirit the rhetoric of God’s conversation interaction within the Trinity and of the Trinity with the created world of words that have being? Could the complex sounds of the semiotic processes of nature be understood theologically as resonances of the divine conversation? Would one find new consonances between theology and the natural sciences in listening and responding to such resonances? From synthetic to polyphonic consonance. Pannenberg’s theology of nature is focused on the problem of how nature’s law-like regularities can be understood in such a sense that they are compatible with God’s action in history, both in particular contingent events and in the final self-­ revelation of God at the end of the history of natural processes, as it is proleptically disclosed in the resurrection of Jesus of Nazareth. Explicitly theological concepts like God’s immensity and God’s eternity provide the framework of possibility in which spaces and times can be understood as forming a unity. The theological perspective always presents the ­anticipation of the totality of being and meaning, which can be fully disclosed only at the end of history. In a way, the understanding of God provides the ground of the possibility of all worldly events in their ­contingent particularity and their emergent connectedness as well as the final horizon in which everything is disclosed in its true relations. ­Theology always offers the synthesis in which everything hangs together. However, everything is based on the “common ground” to which both theology and the natural sciences relate, the relationship between ­contingency and law-like uniformity. The question I would like to raise in conversation with Wolfhart Pannenberg would be whether a polydimensional understanding of reality Martin Luther, Lectures on Genesis 1–5, ed. Jaroslav Pelikan, Luther’s Works (St. Louis, MO: Concordia Publishing House, 1958), 1:21-22. For all fuller exploration of this paradigm shift, see Christoph Schwöbel, “We Are All God’s Vocabulary: The Idea of Creation as a Speech-Act of the Trinitarian God and Its Significance for the Dialogue Between the Sciences and Theology,” in Knowing Creation: Perspectives from Theology, Philosophy and Science, ed. Andrew B. Torrance and Thomas H. McCall (Grand Rapids, MI: Zondervan, 2018), 1:47-69.

32

Wolfhart Pannenberg (1928–2014)

189

would not be more appropriate for the conversations of theology with the natural sciences. Such a polydimensional understanding would comprise at least the physical, the chemical, the biological, the psychological, the personal, the social-historical, the cosmological, and the theological dimensions of reality. Each of these dimensions presents its own order of becoming, its relatedness to the other dimensions, and its own particular modes of communication. On a Christian theological understanding, God is not only the ultimate framework for understanding the different dimensions in their interrelatedness but claims also that this ultimate dimension is disclosed in the dimension of historical and social reality with its physical, chemical, biological, psychological, and personal presuppositions: “The Word became flesh and made his dwelling among us. We have seen his glory” (Jn 1:14). The specifically Christian claim is that the creative and communicative reality is disclosed and becomes accessible in the dimension of a bodily, historical, and social life. And this reality, the eternal creative Logos who communicates by created, historical, and bodily means of sanctification, is disclosed to us through the operation of God’s Spirit. Does this have implications for the way Christian theology conducts the dialogue with the natural sciences? The polydimensional character of reality and the way the triune God is both the ultimate framework and the experiential access to the understanding of reality suggests that the conversation with the natural sciences could be conducted as a conversation with particular sciences without striving in each conversation for the grand synthesis. Would Pannenberg’s engagement with physics appear differently in the context of a more polyphonic conversation with all the sciences? And if the very being of God is a conversation, what better way could there be than being in conversation with particular conversation partners in the sciences? Pannenberg attempts to show the compatibilities of scientific theories and theological reflection. Would we not have to start in our conversations from a lower level and look at the way in which the phenomena are given for scientific investigation and for theological reflection and what they have to say to us when we assume that in the end the phenomena are really legomena of the divine conversation with the created world as it is rooted in the conversation that God is?

190

Christoph Schwöbel

In the end the beginning or in the beginning the end? Another point in the conversation with Pannenberg would have to be his fundamental view that theology looks at all events in the world from the perspective of the end of the history of nature, whereas the sciences proceed by telling the story of the polydimensional becoming of the world from its very first beginnings. Is this the alternative, that theology tries to understand the beginning from the end and the sciences attempt to understand the end from the beginning? It seems that Christian faith as faith in the triune God cannot rest content with this alternative. Is it not committed to the view that if creation is the act of the triune God then it is from the beginning aimed at bringing about the communion of God with his reconciled and perfected creation? It is this end, the end of this beginning, that is anticipated in the operation of the Holy Spirit. What could be gained for the dialogue with the natural sciences if God’s relationship to time is conceived in a consistently trinitarian mode in which God is both eternally present to every event in created time as the creative ground of its possibility and actuality and eternally temporally present to every event in created time and perfectingly present as the consummation of all time in God’s eternity—that is, as the future fulfillment of what everything created is?33 My suggestion to Pannenberg is that only a consistently trinitarian view of the relationship of the eternal God to the times of creation can offer a framework in which God’s eternity is the ground of possibility, the actual fulfillment and the ultimate co-present meaning of each moment, reconciling its past and its future, in the times of creation in the processes of created becoming. On such a view, protology and eschatology cannot be played off against one another. Would this change our view, theologically and perhaps also scientifically, of how contingency and the lawlike uniformities in the processes of nature have to be seen? With this question, the conversation has returned to the initial question with which Pannenberg starts his attempt to explore the consonance between Christian theology and the natural sciences. See Christoph Schwöbel, “The Eternity of the Triune God: Preliminary Considerations on the Relationship Between the Trinity and the Time of Creation,” in Modern Theology 35 (2018): 345-55.

33

Wolfhart Pannenberg (1928–2014)

191

Recommended Reading Primary Works Pannenberg, Wolfhart. Anthropology in Theological Perspective. Translated by Matthew J. O‘Connell. Edinburgh: T&T Clark, 1985. ———. The Historicity of Nature: Essays on Science and Theology. Edited by Niels Henrik Gregersen. West Conshohoken, PA: Templeton Foundation Press, 2008. ———. Metaphysics and the Idea of God. Translated by Philipp Clayton. Edinburgh: T&T Clark, 1988. ———. Revelation as History. Edited by Wolfhart Pannenberg. London: Sheed and Ward, 1969. ———. Systematic Theology. 3 vols. Translated by Geoffrey W. Bromiley. Grand Rapids, MI: Eerdmans, 1992–1998. ———. Towards a Theology of Nature: Essays on Science and Faith. Edited by Ted Peters. Louisville, KY: Westminster/John Knox Press, 1993. ———. What Is Man? Translated by Douane A. Priebe. Philadelphia: Fortress Press, 1971.

Secondary Works Berkson, William. Fields of Force: The Development of a World View from Faraday to Einstein. New York: John Wiley & Sons, 1974. Bradshaw, Timothy. Pannenberg: A Guide for the Perplexed. New York: T&T Clark, 2009. Hesse, Mary B. Forces and Fields: The Concept of Action at a Distance in the History of Physics. Philosophical Library. Edinburgh: Thomas Nelson and Sons, 1961. Reprint, Mineola, NY: Dover, 2005. Schwöbel, Christoph. “Wolfhart Pannenberg.” In The Modern Theologians: An Introduction to Christian Theology in the Twentieth Century. 2nd ed., edited by David F. Ford, 180-208. Oxford: Blackwell, 1997. ———. “The Eternity of the Triune God: Preliminary Considerations on the Relationship Between the Trinity and the Time of Creation.” Modern Theology 35 (2018): 345-55. Thiselton, Anthony C. Understanding Pannenberg: Landmark Theologian of the Twentieth Century. Eugene, OR: Cascade Books, 2018.

9 R O B E RT J E N S O N ( 1 93 0 – 2 0 1 7 ) History’s God Steph en Joh n Wrig ht Robert Jenson was one of America’s most creative and provocative systematic theologians. Although educated in traditional Lutheran contexts— Luther College, Luther Seminary, and the University of Heidelberg—his mature theology developed to become relentlessly ecumenical. While in Heidelberg, Jenson established an enduring friendship and collaborative partnership with Carl Braaten. Jenson and Braaten cofounded two journals (Dialog and Pro Ecclesia) and the Center for Catholic and Evangelical Theology, edited a dogmatics for Lutheran seminary instruction, and produced over a dozen volumes based on the ecumenical meetings of the Center for Catholic and Evangelical Theology. Jenson himself authored over fifteen books and coauthored a volume on Lutheran theology and history. He was also a prolific essayist, with some of his more original ideas peppered throughout journals and edited collections.

S

cripture interrupts the religious desire to deify the world, according to Robert W. Jenson. Where antique religions looked to the semidivine stars in holy awe, the author of Genesis 1 wrote with “deliberate impiety; ‘Gods nothing! Energy sources that God hung up there!’” Jenson wryly concludes, “From here to Galileo is a matter of details.”1 For Jenson, the Christian doctrine of creation follows the lead of Scripture to undermine our more fantastic accounts of the world. Retaining a doctrinal center, Jenson makes frequent and deceptively subtle use of broad scientific insights to develop his theology of creation. Against some recent thought, Jenson does not see science and theology Robert W. Jenson, A Religion Against Itself (Richmond, VA: John Knox, 1967), 22.

1

Robert Jenson (1930–2017)

193

as merely “compatible” or “complimentary” but as both occurring within the one discourse of humanity, making material claims about the world and its significance. Jenson positions the doctrine of creation to fend off two false images of created reality: mechanism and cosmos. God does not create a selfcontained reality, whether imagined as a finely tuned machine that chugs along according to its own internal laws or as a grand cosmos filled with a wondrous spectrum of beings all tending toward transcendence. Both accounts misconstrue the biblical account of creation by failing to account for providence and God’s positive relation to creation. Against the pressure to adopt modern reductionism or to recapitulate antique ontological hierarchy, Jenson proposes something simultaneously more biblical and more fanciful: God creates a history. Science, Modernity, and Theology The question of theology’s relation to the natural sciences arises as a peculiarly modern concern. Jenson learns how to approach modernity from Barth and Jonathan Edwards. Both Barth and Edwards chart a path through modernity, not around it. The reader will find little nostalgia for premodern life in Jenson. Jenson willingly and willfully borrows from the great philosophers of the European establishment, though never simply adopting their methods. Few theologians would attempt to subordinate thinkers as grandiose as Hegel and Kant to their own projects, but Jenson seems to do just that, relishing any moment when he can use their thought against their own purposes. This posture toward modernity is equally evident in his approach toward the natural sciences. Jenson openly accepts the insights of science for theological reflection. He freely speculates on the emergence of life, describes Adam and Eve as the first hominids to pray, and rethinks the doctrine of the ascension in light of the fact that heaven cannot be “up there” somewhere. Jenson’s approach to the sciences is revealed in his judgment of the liberal arts: they are of value because they “complicate us open” (quoting Joseph Sittler). Speaking at Luther College fifty years after the dramas of his early teaching career, Jenson denied that theology

194

Stephen John Wright

should retreat from open engagement with the liberal arts: “Each can open the other to new possibilities of its own power.”2 Such engagement is not without risk. Science has its own explanatory integrity. What theology ought not to do, according to Jenson, is simply fit itself within the story told by science—ultimately capitulating to materialistic reductionism. Closed accounts of scientific judgment do not speak the final word over theological matters. Popular consciousness tends to presume a conflictual relation between science and theology, together offering irreconcilable accounts of the same realities of historical life.3 We can locate here the origins of the modern religious crisis. Is the rainbow a sign of God’s providential care or merely refracted and dispersed light? Equally problematic, however, is the idea that theology and science necessarily provide complimentary but not intersecting modes of discourse. Science can provide epistemological rule over the birds and the seas, and theology will regard the private soul and personal spiritual destiny. For centuries, Jenson laments, Christianity has ceded ground to the sciences, supposing that the two narratives were incommensurable and therefore could be allowed to live in peace only through the compromise of divorce.4 A regrettable early episode saw a division occur between the faculty and administration over a number of issues centering on Jenson, including his openness toward students about evolution. His resignation was refused by the administration, prompting numerous colleagues to offer theirs in protest. 3 A position advanced in the nineteenth century by John William Draper and Andrew Dickson White. Draper’s History of the Conflict Between Science and Religion (New York: D. Appleton and Co., 1874) and White’s A History of the Warfare of Science with Theology in Christendom (New York: D. Appleton and Co., 1896) painted a dire picture of the apparent contest, arguing for the need for the full liberation of science from religion. This position persists because it offers a plausible interpretation of current experience of some of the more extreme religious positions. Recent scholarship has challenged the conflict model. See, e.g., John Hedly Brooke, Science and Religion: Some Historical Perspectives (Cambridge: Cambridge University Press, 1991), who critiques the conflict model for its pretension to reduce all the throes of history to a single conflictual narrative. 4 “For two centuries, we have . . . supposed that Christian talk of God and human destiny must be epistemically disconnected from scientific talk, and that since what science does is describe reality, Christian talk of God and human destiny cannot describe reality. . . . Entire generations and schools of theologians and philosophers have thereupon devoted their careers to inventing some epistemological function for theology.” Robert W. Jenson, “Autobiographical Reflections on the Relation of Theology, Science, and Philosophy; or, You Wonder Where the Body Went,” in Essays in Theology of Culture (Grand Rapids, MI: Eerdmans, 1995), 223. 2

Robert Jenson (1930–2017)

195

Since both of these accounts attempt to locate epistemological boundaries between science and theology, Jenson reasons that they cannot sustain any productive engagement between the two. Any theological account of science that compartmentalizes it within a discrete epistemic realm razes the prospect of human discourse. Science can agitate theology into productive work, Jenson proposes, due to its grounding within the same reality as theology. “If science does not belong to the same discourse as does theology, then science is a play of fictions. . . . No subregion of human discourse can be a normative paradigm of any other, not because they are so discrete but because their boundaries are so blessedly ill-defined.”5 Science can account for the play of immanent causes and realities. From the perspective of theology, scientific description can be understood as the abstracted consideration of the gospel, of the ordering of “God’s history with us.”6 Like Jenson, Rowan Williams does not see the mere “re-enchantment” of nature as a viable option, as it “does less than justice to the scientific enterprise and can all too easily be read as an appeal for less precision—as if it would be better to stop scientific enquiry at a certain stage of complexity or depth.”7 Science as such is in no way unequal to its task—it is a deep and complex investigation of the world of creation. By configuring the relation between science and theology in this way, Jenson allows a “meaningful” scientific narrative to emerge, with neither science nor theology presuming to possess exclusive capacity for the description of reality. The author of Genesis 1, Jenson reasons, deploys the best scientific knowledge available. However, the author utilizes the science in the service of a theology of God and creation. Moreover, he remarks, But even if the science of his day had been altogether wrongheaded, what other science might we have expected him to deploy? The science of a millennium or so after his time, which would have been gibberish for him and his readers? Or a revealed and thus timelessly satisfactory science, which would have been no science at all?8 Jenson, “Autobiographical Reflections,” 224. Robert W. Jenson, Systematic Theology (New York: Oxford University Press, 1997–1999), 2:45. 7 Rowan Williams, The Edge of Words (London: Bloomsbury, 2014), 120. 8 Robert W. Jenson, Canon and Creed (Louisville, KY: Westminster John Knox, 2010), 98. 5 6

196

Stephen John Wright

The findings of science are always provisional. Since science is methodologically open to revision and the displacement of theories, its findings cannot be determined in advance by any passage of Scripture. We do not here rule out consonance between science and the scriptural record, but it is imperative to the scientific task that no accord be presumed. While a timid theology might want to avoid the risk of taking on the provisional findings of current science, liable as they are to change, Jenson sees no such hesitation within Scripture. He claims to model his interaction with provisional scientific claims on the “priestly savants who constructed Genesis 1, who took the risk and came out pretty well.”9 Creation History Jenson admires the speculative boldness of the early Christians who threw off the hierarchical Platonic ontologies in an attempt to take seriously Scripture’s presentation of reality. The picture of reality given in Scripture involves only two kinds of being: the Creator and the creatures. Most basically, we may say that “the Creator is one who does something, and . . . creatures are what he does.”10 This relation cannot be reduced, Jenson reasons, to other conceptual pairings such as infinite and finite or immanent and transcendent. Such pairings are useful but only because they gather their meaning from the fundamental relation of the Creator to creatures. Jenson argues that “in scripture itself, the difference between the Creator and his creatures is not laid out conceptually at all, but rather narratively.”11 The two do not form a neat conceptual—or even dialectical —couple that can be described through various adjectival pairings. Theology evokes the relation of God to creation only by commenting on the story of Scripture. That Scripture’s way of describing creation runs through narrative is an idea that Jenson finds endlessly provocative. Speaking of Genesis, Jenson argues that there is no reason why “a poem about hearing creation Jenson, Systematic Theology, 2:43n64. Robert W. Jenson, “Creator and Creature,” International Journal of Systematic Theology 4, no. 2 (2002): 217. 11 Jenson, “Creator and Creature,” 219. 9

10

Robert Jenson (1930–2017)

197

cannot be serious ontology.”12 God speaks creatures into being, and we arise as the creatures God narrates. In this, Jenson finds a mandate to avoid all ahistorical attempts to account for creation. God, he proposes, creates a history. Few prove as important for Jenson’s doctrine of creation as Jonathan Edwards. Jenson discovered Edwards midway through his career. Essentially a figure of the Enlightenment, Edwards used Locke and Newton to new ends. These developments were not immediately seen as good news for Christianity. Newton’s mathematical description of physical motion made possible the account of the world as a play of purely immanent causes. With Newton and the Enlightenment, the old Aristotelian causes began to drop away until only efficient causality was left. A world imagined as animated solely by immanent efficient causality would be something like a machine. Jenson found himself drawn to Edwards’s particular method of denying mechanism.13 Rather than rejecting Newton’s physics, Edwards put Newton to metaphysical use. The problem, after all, was not with his description of motion but rather the metaphysical implications that people chose to draw from that description. Edwards, according to Jenson, did not dispute the scientific framework. Instead, he diagnosed an unexamined conflation of Newton’s “bodies” with “substance,” understood as the “supposed hidden, intrinsically potent subject of . . . overt attributes and actions.”14 If bodies are substances in this sense, then the mechanism metaphor has purchase, but Edwards set out to deny that bodies had either intrinsic subsistence or potency.15 God alone, according to this stipulation, could be substance.16 “Substance,” according to Jenson, “is a God-concept.”17 It envisages a reality of self-established integrity. The temptation in Edwards’s time was to read the emerging theories on atoms as accounts of substance. Edwards refused to credit Jenson, Systematic Theology, 2:157. As found, e.g., in Jonathan Edwards, The Works of Jonathan Edwards, vol. 6, Scientific and Philosophical Writings, ed. Wallace E. Anderson (New Haven, CT: Yale University Press, 1980), 216, 378. 14 Robert W. Jenson, America’s Theologian (New York: Oxford University Press, 1988), 25. 15 For instance, in Edwards, Works of Jonathan Edwards, 6:215. 16 And even in this case, only loosely. Jenson, America’s Theologian, 26. 17 Jenson, America’s Theologian, 26. 12

13

198

Stephen John Wright

that any scientifically observable phenomena could be considered “substance.” In Edwards’s view, the notion of substance has a certain obvious though deceptive logic to it: “It is exceedingly natural to men to suppose that there is some latent substance, or something that is altogether hid, that upholds the properties of bodies . . . men are wont to content themselves in saying merely that it is something; but that ‘something’ is he by whom all things consist.”18 In Jenson’s summary, there are no “little selfsufficient agencies beside God, natural entities are not godlets, and therefore the world harmony is not self-contained.”19 In making this argument, Edwards posited immediate divine agency as the basis for atomic integrity.20 Such a view cannot be sustained by the standards of today’s physics, which no longer understands atoms as indivisible units of matter. The old atomic model that provided the basis for Edwards’s speculation no longer has any hold in the sciences. However, Jenson does not see this as a weakness in Edwards’s argument. Since the motivation behind his arguments was the dismissal of substance, Jenson sees the dismissal of classical atomic theories as a vindication of Edwards’s basic insight: “Physics now posits no entities that satisfy the concept of substance, just as Edwards insisted it should not.”21 Troubling the notion of substance within the material created order significantly weakens the metaphor of mechanism, as there are no self-existent realities to be assembled into a mechanism. The physical and atomic models do not trouble the metaphysical point that Edwards was laboring over. The concept of substance, as Edwards saw it used in his time, enabled deistic renderings of the universe. Any metaphysic that speaks of self-contained integrities naturally brackets out God. This was Edwards’s point. Although the scientific features of his argument no longer hold, Jenson argues that the metaphysical arguments continue to have relevance. Christian engagement Edwards, Works of Jonathan Edwards, 6:380. Jenson, America’s Theologian, 26. 20 “We may speak in the old way, and as properly and truly as ever: God in the beginning created such a certain number of atoms, of such a determinate bulk and figure, which they yet maintain and always will; and gave them such a motion, of such a direction, and of such a velocity. . . . Yet perhaps all this does not exist anywhere perfectly but in the divine mind.” Edwards, Works of Jonathan Edwards, 6:354. 21 Jenson, America’s Theologian, 29, emphasis in original. 18 19

Robert Jenson (1930–2017)

199

with the sciences must be alert to the ways in which root metaphysical presuppositions can undermine basic Christian claims. Adapting Edwards’s insights into his own system, Jenson further points to the ahistorical character of a mechanistic universe. The basic character of such an account of creation would be the sheer “thereness” of creation, with no thought as to intention and direction. Building a theology of creation from the Creator/creature distinction, however, does not accord us the liberty to think of the world as sheer factual matter. Rather, the Creator God of the Christian gospel relates to the world through covenants and promises, expressing intention not just that things should exist but that they should be caught up in the dramatic turns of history. Like the mechanism metaphor, the ancient depiction of creation as a “cosmos” sometimes picked up by theologians suggests that God’s creative act instigates, but does not guide, the direction of the world. “The world God creates is not a thing, a ‘cosmos,’ but rather is a history. . . . The call of Abraham, the Exodus, the Crucifixion and Resurrection and the final Judgment are not events within a creation that is as such ahistorical; they are events of the history that is created.”22 The Spirit of History Denying mechanism and cosmos in this way calls into question the notion of “divine intervention” in history. The metaphor of intervention is predicated on the idea that the universe is self-contained and would otherwise simply run its course. A world-machine does not need a creator, only an engineer. And the engineer’s agency regarding a mechanism can only be thought of as “intervention . . . fooling around with monkey wrenches . . . fixing what ain’t broke.”23 As Jenson observes, the standard accounts that employ these metaphors of intervention struggle to make sense of God’s movements throughout history. Under the conditions of materialism, the direction that history actualizes out of the various possibilities available to it can be only either determined or random. Refusing both of these options, Jenson, Systematic Theology, 2:14. Jenson, “Autobiographical Reflections,” 217.

22 23

200

Stephen John Wright

Jenson argues that history’s sojourn demonstrates a freedom and “spontaneity” identified with the work of the Spirit.24 God has no need of intervening in history, like a mechanic tuning an engine, since creation has no direction of travel apart from the freedom of the Spirit’s agency. Lest anyone think that Jenson posits here a “God of the gaps,” he clarifies, “We have not located the liberating agency of Christ’s Spirit in regions supposedly not covered by scientific description; what is attributed to the Spirit is a universal feature of the world precisely as scientifically described.”25 Through this pneumatology, Jenson appears to be evoking a depiction of the God-world relation somewhat akin to Aquinas’s distinction between primary and secondary causality. It makes sense to pray, Jenson argues, because the “scientifically accountable actual course of events can and so must theologically be understood as a history occurring within God’s Freedom.”26 Divine responses to prayer are not “interventions” that break or suspend the laws of the material universe, since those laws have no reality or direction apart from the free act of the Spirit. Jenson distances this view from panentheism, since the “in/out” distinction has been rejected by the apprehension that “God wills to know a world, and this world rather than some other; thus the world is willed reality and God is reality that does not need to be willed.”27 It would thus be inappropriate to measure God’s relation to creation by distance or direction. Has Jenson, then, summoned God into time by appealing to the Spirit in this way? Here we find Jenson’s most novel and radical departure from the tradition. Jenson refuses to define eternity as timelessness, avoiding equivocation in this act. As indicated earlier, Jenson grounds all distinctions between God and us on the fundamental distinction between Creator and creature. Thomas Aquinas’s doctrine of analogy functions only because of the reality of this unique distinction. Time itself, Jenson proposes, must be analogous. Given analogy, we can describe God’s relation to time positively, not by sheer negation. He cites this analogical Jenson finds support for this view and its consonance with current scientific opinion from a paper by George Murphy, “The Third Article in the Science–Theology Dialogue,” Perspectives on Science and Christian Faith 45, no. 3 (September 1993): 162-68. 25 Jenson, Systematic Theology, 2:43, emphasis in original. 26 Jenson, Systematic Theology, 2:44. 27 Robert W. Jenson, On Thinking the Human (Grand Rapids, MI: Eerdmans, 2003), 52. 24

Robert Jenson (1930–2017)

201

treatment of time as the main differentiating feature between Aquinas’s thought and his own.28 Moreover, analogy upholds the fundamental distinction between Creator and creature, avoiding melodramatic claims of God’s susceptibility to the travails of history. With Augustine, Jenson treats time as a creature; like all creatures, time belongs to the providential care of God. Providence, here, signals God’s intimate presence to all of time—past, present, and future. Time Whereas Jenson finds himself to be ambivalent about the traditional language of being, he sees as utterly indispensable the language of time. The God revealed in the narrative of Scripture neither rescues us from time nor poses time as an enemy to be overcome. Scripture instead tells of God’s triumph through time. Whatever faithlessness of the chosen people, God’s promises will be upheld. Whatever tragedy befalls the incarnate Son, God’s life will shine forth. In Jenson’s discussions of time, he regularly draws a parallel between the philosophical arguments about time and Stephen Hawking’s description of the distinction between “imaginary time” and “real time.” Hawking’s scientific argument, in Jenson’s judgment, is fundamentally “theological.”29 The distinction entails nothing less than fundamental cosmogony. Real time, if it has motion, always moves in one direction. One might think of Newton’s “absolute time” here, but Hawking has something more like time according to relativity in mind.30 The universe, according to real time, may have a beginning and an ending.31 However, on the quantum model that Hawking investigates, the notion of imaginary time might account for the universe without a linear temporality. Imaginary time is here understood according to a spatial metaphor, sitting at a right angle to the continuum of real time. As Jenson observes, imaginary time Jenson, Systematic Theology, 1:55n87. Robert W. Jenson, “Does God Have Time?,” in Essays in Theology of Culture (Grand Rapids, MI: Eerdmans, 1995), 196. 30 Jenson notes the connection between relativity and real time in Hawking. Jenson, “Does God Have Time?,” 197. 31 As Hawking puts it so plainly, according to real time it makes sense that we remember the past but not the future. Stephen Hawking, A Brief History of Time, rev. ed. (New York: Bantam Books, 1996), 148. 28 29

202

Stephen John Wright

allows for the arrows of time to become reversible.32 Or, perhaps, we might think of imaginary time as a conception of time not determined by arrows and direction. Quantum thought supposes that time functions rather like the physical dimensions, extending in space. By reading Hawking in this way, Jenson wishes to draw a comparison to usual theological conceptions of time. “Which is the really real time?”33 Shall we account for time by appeal to an account of reality in which there are no beginnings or endings? Hawking’s preference for imaginary time and its potential for an eternal—timeless?—universe Jenson interprets as capitulation to the old religious security of a timeless eternity in which past, present, and future have no meaning. Imaginary time, as much as timeless eternity, Jenson regards as the negation of the intelligibility of life. Jenson argues that something of both real time and imaginary time must hold if time is, as Kant suggests, the “horizon of experience.”34 Real time accounts for time as the context of our experience. Jenson argues, inasmuch as we experience anything, we experience time. He associates this internal experience of time with Augustine. Imaginary time, by contrast, posits that time functions as another spatial axis—a literal “fourth dimension.” This provides the external structure for reality of time, allowing time to be the horizon of experience. Imaginary time, in its externality, models the Aristotelian tradition of time. As such, Jenson sees in Hawking the outworking of a familiar metaphysical problem: Shall we think of time as inner experience or as the external world-structure? Hawking attempts to push this question back onto the equations to let science decide for us, but Jenson holds that Hawking has played his cards already—he has his own metaphysical ambitions and gives preference to imaginary time precisely because it problematizes theologies of creation.35 Mapping the same distinction onto space, Thomas F. Torrance has argued that the theologians of the Nicene era denied the Aristotelian model of space as a container grounded by an immobile center. Rather, the “sensible world (kosmos aisthetos) and the intelligible world (kosmos Jenson, “Does God Have Time?,” 197. Toward the end of his life, Hawking proposed that traveling backward in time was a physical impossibility. 33 Jenson, “Does God Have Time?,” 197. 34 Jenson, “Does God Have Time?,” 198, emphasis in original. 35 Jenson, Systematic Theology, 2:34n20. 32

Robert Jenson (1930–2017)

203

noetos) . . . came to be understood . . . as actually intersecting in Jesus Christ.”36 God’s eternity, for Jenson, can be neither mapped onto a timeline nor construed as a “timeless point from which all points on the time line are equidistant.”37 Instead, time and eternity meet in Christ. Such an account does not entail paradox—a timelessly temporal incarnate Christ—but an analogous complementarity. Time, in Jenson’s theology, “is the room God makes in his eternity for others than himself.”38 For a doctrine of creation to function coherently, Jenson offers, it must be founded in the primacy of time. “God does not create spatial objects that thereupon move through time; he creates temporal-spatial objects, that is, in a more precise language, he creates histories.”39 One implication of this commitment is a growing wariness about Hawking’s proposals that would potentially collapse time and space. If imaginary time models time as another spatial axis, what reality does time have in distinction from space? Jenson proposes that this is a relation that Christianity—indeed, humanity—cannot do without. At stake in time’s priority to space is nothing less than the coherence of human experience. Were imaginary time to eclipse real time as the “really real time,” human experience of time’s flows and tensions would be rendered a fanciful illusion. And the treatment of our experience of time as an illusion to be overcome is nothing less than the model of religion that Christianity sought to overcome, in Jenson’s view. As Jenson attempts to show, a denial of a negative relation between God’s eternity and creaturely time need not result in having God enter into time and undergo it like a creature. One of his strongest statements on this appears in a unique book: a transcript of conversations he held with his then eight-year-old granddaughter, Solveig. When discussing how time and eternity relate, Jenson says, “God doesn’t have a timeline. . . . That’s the difference between us.” When pressed by Solveig, “But how do you imagine eternity?” Jenson responds flatly, “You don’t imagine it.”40 Thomas F. Torrance, Space, Time and Incarnation (Edinburgh: T&T Clark, 1997), 15. Jenson, Systematic Theology, 1:140. 38 Jenson, Systematic Theology, 2:46. 39 Jenson, Systematic Theology, 2:46. 40 Robert W. Jenson and Solveig Lucia Gold, Conversations with Poppi About God (Grand Rapids, MI: Brazos, 2006), 143. 36 37

204

Stephen John Wright

Being Human Theological accounts of human uniqueness center on the concept of the imago Dei. This alone seems to distinguish humans from other animals, who all are nephesh, according to the author of Genesis. Factoring in the insights of evolution, we might say that this is a feature that humans have but our prehuman ancestors presumably did not. The imago Dei, therefore, is acquired or emerges over time. Jenson freely ruminates on such ideas in his account of theological anthropology. Elements of Darwinian evolution appear in this account, utilized for theological ends.41 The Darwinian concept of adaptation provides the background to his account of the first humans. Speaking at a symposium titled “Ritual and Humanizing Adaptation,” Jenson pressed the premise of the discussion as a theologian: “‘Adaptation’ to what? If our culture’s standard association of terms is to be followed, ‘adaptation’ is to the ‘environment,’ a term devised on purpose to bracket out the reality of God.”42 If it is to avoid base materialism, a theological account of the human will raise the question of adaptation to God. Jenson moves here from sheer biological adaptation—whatever constitutes humanity as a species does not capture what Christianity says about humans as creatures—to consider the way that humans are suited to be the objects of God’s redeeming action. Jenson supposes that the joining of evolutionary theory and theology might prove to be fruitful in considering humanity. However, there are some obstacles to be overcome. Adaptation as an evolutionary function does not describe a mechanism so much as offer a descriptive account for the success of chance mutations in allowing organisms to survive and flourish within a given environment. Or such was Darwin’s intuition: “I am inclined to look at everything as resulting from designed laws, with the details, whether good or bad, left the working out of what we may call chance.”43 Adaptation, in this sense, finds its determination in a reality already given. Jenson’s early work gives no sources for his understanding of Darwin. He appears to work with broad and general renderings of the theory. Later, in his Systematic Theology, he directs readers to Richard Dawkins and a piece by Christof K. Briebricher, a German biophysical chemist who contributed to a volume on science and theology, for a survey of the field of evolutionary knowledge. Jenson, Systematic Theology, 2:127-28. 42 Robert W. Jenson, “The Praying Animal,” in Essays in Theology of Culture, 117. 43 Charles Darwin, “Letter to Asa Gray,” in Charles Darwin, ed. Francis Darwin (London: John Murray, 1908), 236. 41

Robert Jenson (1930–2017)

205

For Jenson, this account cannot quite work for Christian theology. Whatever adaptation might occur through Christian ritual, it is not the adaptation to a given—to a God whose being is established from the past. The reason for this is evangelical: Christianity takes its orientation from the death and resurrection of Jesus, “the most radical possible disruption of continuity and development.”44 Christ’s resurrection poses to us not the securities of the past but the openness of the future. Christian spirituality is built on the possibilities occupying the horizon of experience, given that Jesus lives. Wrenching the terms of the discussion somewhat, Jenson proposes that theology instead consider what it would mean to adapt to a God from the future. A theology so shaped will not yet be done with teleology. While modern sciences tend to avoid teleological accounts—most of which have the appearance of covert theological commitments—Jenson asks, “Why should what an event or condition comes from explain it more appropriately than does what comes of that event or condition?”45 To suppose that origins alone determine the meaning of the world reveals a metaphysical commitment incompatible with the gospel. Without weighing in on the arguments about the anthropic principle, Jenson holds that it is not insignificant that humanity exists in the universe as it is, “black holes and variously flavored quarks and free-floating bits of genetic information” and all.46 Perhaps the rate of the expansion of the universe has meaning precisely because it allowed the perfect conditions for carbon-based life at this point in its movement. Similarly, Jenson holds that human life has its meaning in what it will be rather than from its origin. For Jenson, the apex of human adaptation arrives in the act of prayer. In prayer, humans orient themselves toward the future promised by God. Prayer provides the basic theological definition of Jenson, “Praying Animal,” 118. Jenson, “Praying Animal,” 130. Jenson continues on this line in his systematics: “The exchange ‘What big teeth you have, Grandma!’ ‘The better to eat you with, my dear!’ is indispensable in the story and therefore far closer to reality than any nonteleological reduction can be.” Jenson, Systematic Theology, 2:128. 46 Jenson, Systematic Theology, 2:116. Jenson’s knowledge of the anthropic principle comes from his reading of John D. Barrow and Frank J. Tipler, The Anthropic Cosmological Principle (New York: Oxford University Press, 1986). 44 45

206

Stephen John Wright

the human: “Who then were Adam and Eve? They were the first hominid group that in whatever form of religion or language used some expression that we might translate ‘God,’ as a vocative.”47 The act of prayer establishes humanity without regard to biological features. “Theology need not share the anxious effort to stipulate morphological marks that distinguish prehumans from humans in the evolutionary succession.”48 The first prayer that establishes our species as human need not have coincided with the appearance of the biological markers that constitute our species.49 Answered Prayer What Jenson admires about Edwards in particular is that his rejection of a mechanistic universe did not signal a retreat into an anterior classical universe but rather a striking out into new speculation about a modern scientific universe. Edwards’s venturesome theology showed that the problems thrown up by science frequently rearticulate problems already internal to theology. By locating the disagreement between science and theology within metaphysics, Jenson frees up both to their own tasks. “Each can open the other to new possibilities of its own power.” Jenson resists modernity by way of modernity. The demythologizing of the world is a requisite both of science and theology: “So long as any aspect of the world is divine, science cannot start. We do not shoot rockets at Venus if we think it the goddess of love.”50 Positively, Jenson views the undertaking of science as a commitment to the meaningfulness of reality. To investigate the world, one must trust that the world is filled with meaningful relations. AnyJenson, Systematic Theology, 2:59. Jenson, Systematic Theology, 2:59. 49 Science has indeed moved on since Jenson wrote this, with the restructuring of the evolutionary taxonomy to create space for hominins, a category inclusive of all the homo species, including humanity. To use hominid as Jenson does would be imprecise today. Eugene Rogers notes this shift and sees in it the risk that theology undertakes when it internalizes scientific knowledge. Despite the danger, he judges that Jenson’s theology does not suffer the change, since Jenson does not attempt to use theology to replicate the insights of science. Eugene F. Rogers Jr., “The Blood of Christ and the Christology of Things: Or, Why Things Became Human,” in The Promise of Robert W. Jenson’s Theology: Constructive Engagements, ed. Stephen John Wright and Chris E. W. Green (Minneapolis: Fortress Press, 2017), 159. 50 Robert W. Jenson, Story and Promise (Philadelphia: Fortress Press, 1973), 152. 47 48

Robert Jenson (1930–2017)

207

thing less would not compel one to “investigate the heredity of a flea with the same enthusiasm as the heredity of an emperor.”51 Theologians have little call to quibble with science’s descriptive account of creation, but theology’s task may lie in interrogating the accompanying metaphysics. By shelving the metaphors of mechanism and intervention, Jenson evokes a different metaphysical landscape on which to imagine the relation of science to theology: history. Such an account opens up space for thinking providence. We have no reason to suppose that the question of providence will not become more urgent in the church in coming years, as petitionary prayer and charism become increasingly central to the growing forms of Christianity in the global church—or as political and ecological arrangements demonstrate a sparsity of tangible beneficence. Few theologies seem equipped to account for answered prayer. Prayer, Jenson has regularly suggested, must be understood as involvement in providence. “If, of course, God does not in fact freely manage the universe, petitionary prayer can’t work—but then neither is any other form of Christian prayer meaningful, for then there isn’t the Christian God.”52 But what does one expect when praying for Margot to recover from her illness, or to pray that the world might prosper despite human-inflicted damage to its vital systems? Nothing more, Jenson seems to suggest, than that God be the God of history, the narrator of our common tale. Under these conditions, we can no longer think of intervention. After all, narrators do not intervene in stories; they tell them. An answered prayer could be only a moment of the Spirit’s freedom in the movements of God’s history, whatever a scientist might see. Jenson, Story and Promise, 152. Robert W. Jenson, “Appeal to the Person of the Future,” in The Futurist Option, ed. Carl E. Braaten and Robert W. Jenson (New York: Newman Press, 1970), 157. Jenson notes the work of Frank Tipler and George Murphy to suggest that a certain reading of the anthropic principle would enable something like a Christian eschatology, making resurrection a necessary outcome of the laws of physics. Rather than investing heavily in these controversial ideas, Jenson posits the freedom of the Spirit as accounting for natural events and the direction of history. One guesses that he sees Tipler’s science as offering something too close to a mechanism for resurrection. Jenson, Systematic Theology, 2:42n62. See George L. Murphy, “The End of History in the Middle: Speculation on the Resurrection,” in Works, vol. 5:2 (1995), 1–3 (Chicago: ELCA Work Group on Science and Technology); and Frank Tipler, The Physics of Immortality (New York: Doubleday, 1994).

51

52

208

Stephen John Wright

Capacious Doctrine A popular trope suggests that theology deals in eternal realities untouched by time, whereas science investigates the rougher contingent structures of embodied life. We have already seen how Jenson transgresses this artificial boundary. Upon closer examination, we begin to see the limits that Jenson places on science’s role in theology. Jenson works with what appear to him to be agreements between theology and science. But he does, in one work, caution against making too much of these agreements, “since even the most firmly established structures of science are proper science only in that they remain open to being displaced.”53 One must take care not to overstate the finality of present scientific consensus, but here emerges a correlate problem. Just how might one go about holding scientific insights lightly without needlessly diminishing science’s capacity to render true insights into the world? The problem troubles both science and theology. How are they to remain themselves in this engagement? To answer this, we might look to Jenson’s understanding of the task of theology. Jenson understands theology as second-order reflection on the Christian task of speaking the gospel. The chief theological question, therefore, amounts to “What must one be saying to be saying the gospel?”54 The question entails cultural and historical contingency. For Jenson, this question must be asked anew at all times; in this conception he unsettles the finality of theology’s insights. Theology, in this conception, cannot be reduced to mere reiteration of revealed timeless truths but is a critical task aimed at the proclamation of the Christian gospel. This critical task “can turn to anything at all.”55 Jenson models a mode of engagement with the sciences not engaged primarily in apologetics but with the more fundamental project of proclamation: lex proclamandi lex credendi. He thus draws out resonances that sound between theology and science, co-opting the insights of science to further the task of theology. Jenson, Canon and Creed, 97. Jenson, Systematic Theology, 1:3-22. 55 Jenson, Systematic Theology, 1:20-21. 53 54

Robert Jenson (1930–2017)

209

How can theology survive the shifting understandings of science? Perhaps, to remain true to its task, theology ought not to aim for a fully scientific theology.56 We can be instructed on this point by Jenson’s engagements with science more than by his statements about science. I propose that Jenson theologizes in a way that accommodates science without being directed by it. Jenson accommodates science by bringing into alignment scientific observations with basic Christian claims that can hold independent of this alignment. When Jenson says, “Who were Adam and Eve? They were the first hominid group that in whatever form of religion or language used some expression that we might translate ‘God,’ as a vocative,”57 he has not created a radically new theological anthropology. With a few terminological substitutions, the six-day creationist can make the same theological confession. Instead, Jenson has articulated the theology of the tradition in a way that is open to current scientific findings. In the same way, he proposes that one can continue to believe in heaven after Copernicus by drawing on the riches of Lutheran Christology, concluding that God does not need a space to occupy, since God is God’s own space.58 Heaven is wherever God is. These pithy aphorisms open outward to create space for engagement with science. However, they continue to stand as theological teaching apart from any given particular scientific claims. We can see this pattern of aphoristic accommodation most clearly in his gloss on Genesis: God creates by commanding; the existence of the world is an act of obedience to his command; the initial result of his communication is an explosion of energy. . . . “Light” is of course Genesis’s word for what we would now call “energy.” . . . That the universe begins as a burst of sheer energy, that is then endlessly differentiated, is a familiar story. . . . Moreover, we might go on to note that the Genesis story of cosmic differentiation is followed by a narrative of chronologically successive life forms, moving from the relatively simple to the more complex—another familiar story.59 One need only to substitute a few terms in Balthasar’s critique of “aesthetic theology” to see the peril here. 57 Jenson, Systematic Theology, 2:59. 58 “Christ has risen to be in God’s place. God, however, is in no place but is his own place; and over against God, the created universe is therefore just one other single place.” Jenson, Systematic Theology, 203-4. 59 Jenson, Canon and Creed, 94, 97. 56

210

Stephen John Wright

By a suggestive formal comparison, Jenson accommodates the broad contours of contemporary cosmology. By appeal to structural similarity, Jenson’s theological provocations are unlikely to be entirely unseated by new scientific insights, grounded as they are in established doctrine and Scripture rather than contingent science. One final observation needs to be made on this front. Jenson maintains the traditional distinction between doctrine and theology. Engagement with the sciences operates at the level of theology, not doctrine. As doctrine is the teaching of the church and no single individual, no single theologian can by their experiments with science undermine doctrine. Far from dealing in timeless truths, Jenson considers theology to be historically conditioned and liable to revision. Each theological work “dismembers its predecessors and uses the fragments in strange ways.”60 As such, even if new science caused Jenson’s theology to implode, he would shrug and concede, “that is how it goes when mere humans try to do theology.”61 Conclusion Engagement with the sciences never featured prominently in Jenson’s theology.62 More than anything else, he was a dogmatician committed to calling the church back to its doctrinal center. Nevertheless, there are some lessons to receive from Jenson’s confident use of scientific insights. His refusal to locate the theological and the scientific within different epistemic realms creates space for the free exploration of theological ideas through the sciences. Jenson models generalist engagement—an approach to the sciences that bypasses the usual science-religion apologetics and the need for dual doctoral qualifications in science and theology. One might reasonably expect that science’s methodological convictions place it necessarily at odds with Christian theology, but Jenson sees no fatal vulnerability here.

Jenson, Systematic Theology, 2:vi. Jenson, Systematic Theology, 2:vi. 62 His awareness of the sciences tended to arise through the work of celebrity scientists—Stephen Hawking, Richard Dawkins, etc.—or from other theological engagements with science. 60 61

Robert Jenson (1930–2017)

211

The secularization of nature allowed by Christian Scripture63—“When you look up to the heavens and see the sun, the moon, and the stars, all the host of heaven, do not be led astray and bow down to them and serve them” (Deut 4:19 NRSV)—enables free scientific advancement. Jenson finds nothing theologically problematic in living according to the scientific “ethos”: “We can allow no barriers to our investigation: we must see the very plasm of life as a replicable reaction-system, and tramp across the face of the moon, as soon as we can make a way to get there.”64 The Christian doctrine of creation, in Jenson’s view, enables science by freeing us from the worship of the world. What can go wrong with science, however, is that it succumbs to nihilism, treating the universe as meaningless chaos rather than a field of wonders. Only when secularization becomes divorced from the gospel does creation become ­susceptible to exploitation, and science can function destructively: ­liberation becomes license.65 Jenson thereby provides us with a model for a somewhat low-key engagement with science. He articulates doctrine in a way that accommodates scientific insight. This account differs from the “compatibility” accounts, which suggest an epistemic division between theology and science. Instead, Jenson allows us to imagine what theology might be like were theologians not to foreclose the possibility that science and theology speak of the same reality: the world in Christ. Recommended Reading Primary Works Jenson, Robert W. Story and Promise. Philadelphia: Fortress Press, 1973. ———. Systematic Theology. 2 vols. New York: Oxford University Press, 1997–1999. ———. Theology as Revisionary Metaphysics: Essays on God and Creation. Edited by Stephen John Wright. Eugene, OR: Cascade Books, 2014. ———. A Theology in Outline: Can These Bones Live? Edited by Adam Eitel. New York: Oxford University Press, 2016. Jenson acknowledges Lynn White’s influential critique of Christianity but does not directly disagree with it. He appears to offer tacit agreement with the basic history but disagrees with the conclusions. Jenson, Systematic Theology, 2:114n13. 64 Jenson, Story and Promise, 153. 65 Jenson, Systematic Theology, 2:115. 63

212

Stephen John Wright

Secondary Works Green, Chris E. W. The End Is Music: A Companion to Robert W. Jenson’s Theology. Eugene, OR: Cascade Books, 2018. Wright, Stephen John, and Chris E. W. Green, The Promise of Robert W. Jenson’s Theology: Constructive Engagements. Minneapolis: Fortress Press, 2017.

10 C O L I N E . G U N T O N ( 19 4 1 – 2 0 0 3 ) The Triune God, Scientific Endeavor, and God’s Creation Project Mu rray A. R ae Colin Gunton was born in 1941 and died in 2003. He was an English theologian who spent the whole of his teaching career at King’s College in London. His many books and articles variously explore the subject matter of Christian doctrine, but all are guided by his insistence that the being and action of the triune God provides the decisive clue to the essential nature of all reality and to all that is going on in the world. He brought this conviction to bear on a reformulation of the doctrine of creation and to a profound analysis of human culture. Gunton’s work attracted graduate students from around the world to the Research Institute in Systematic Theology, which he cofounded in 1988 with Christoph Schwöbel. In 1999 Gunton, along with John Webster and Ralph Del Colle, established the International Journal of Systematic Theology, which quickly became one of the leading journals in the field.

O

ne of the enduring contributions of Colin Gunton’s theological career is his development of a robust account of the doctrine of creation for the contemporary world. Gunton considered it necessary to reestablish the doctrine on solid trinitarian grounds and to free the doctrine from what he saw as the wrong kind of entanglement with modern science. He believed that theology had become beholden to modern science in ways that adversely affected the proper investigation of its own subject matter—namely, the being of the triune God and God’s dealings with the world. He was also convinced that science and theology can become fruitful partners in our efforts to understand the nature of reality only when each discipline is free to pursue its own investigations in ways

214

Murray A. Rae

appropriate to its own subject matter. Efforts to resist (in the name of theology) scientific investigation of the evolutionary adaptation of species, for example, are as misguided as attempts made in the name of science to resist the biblical claim that God created and sustains the world. Colin Gunton was raised within the Congregational Church rather than the Church of England. This detail is important, for Gunton often referred to himself as a dissenter. He was a dissenter from the notion that the church should be tied constitutionally to the state, but he also dissented from the prevailing disposition in English theology during the time of his theological education at Oxford, in which theology had been subjected, more or less willingly, to norms of thought drawn from outside its own proper calling and sphere of responsibility. In the 1960s, those norms were drawn especially from science and from a particular conception of science determined by post-Enlightenment rationalism. Theology, it was thought, must conform to the conceptions of truth and knowledge delivered by the Enlightenment, especially to the view that through the proper exercise of human reason there could be obtained an objective and impartial grasp of the nature of reality.1 This “modernist heresy,” as Gunton called it, issued in the quest for “a purely secular foundation for knowledge [that] amounted to a human attempt to displace God as the source of being, meaning and truth.”2 The widespread acceptance of these constraints among English theologians of the mid-twentieth century resulted not only in a false conception of the authority of science but also in a loss of proper confidence in the foundations and central content of Christian faith— namely, the incarnation, life, death, resurrection, and ascension of Jesus. According to Gunton, it is this historically contingent reality, given by God and made accessible to us through the work of the Holy Spirit, that “determines the very essence of Christian faith”3 and constitutes the subject matter to which Christian theology must attend. Attention to this subject matter also yields a theological conception of On which, see Colin E. Gunton, Enlightenment and Alienation: An Essay Toward a Trinitarian Theology (London: Marshall, Morgan & Scott, 1985; repr., Eugene, OR: Wipf & Stock, 2006). 2 Colin E. Gunton, “Theology in Communion,” in Shaping a Theological Mind: Theological Context and Methodology, ed. Darren C. Marks (New York: Routledge, 2017), 33. 3 Gunton, “Theology in Communion,” 31. 1

Colin E. Gunton (1941–2003)

215

human culture more broadly, including the activity we call science. The legitimacy, the value, and the success of science, according to Gunton, is best understood within a doctrine of creation framed and determined by the activity of God. Undaunted by opposition to this resolutely theological understanding of the world, Gunton devoted himself to the pursuit of a theology determined in content and method by the self-disclosure of the triune God. His theological career may be characterized, I suggest, as the joyful pursuit of the possibilities that emerge for life and for thought through faithful attention to this subject matter. While Gunton, following in the wake of theologians such as Karl Barth and T. F. Torrance, was slowly freeing himself from the grip of the Enlightenment and the undue constraints that it put on theology, we should not too quickly assume that he rejected all the historical, cultural, and intellectual developments surrounding it. While, in Gunton’s judgment, the Enlightenment led us down some false paths—most notably in its displacement of God, in its epistemology, and in its problematic conception of human freedom—he was nevertheless sympathetic to some of the criticisms it offered of traditional Christianity. “The Enlightenment’s view of traditional Christianity as authoritarian and excessively other-worldly was not entirely a caricature,” he writes, while he finds common cause with Enlightenment thinkers who reject “hierarchical” ontologies and the absolutist political regimes that such ontologies commonly spawn.4 The challenge, for Gunton, lay in figuring out where our modern culture is a proper development of the Christian tradition, where it is reacting against improper excesses of the tradition, and where it is a deviation from Christianity’s basic theological foundations. More than an exercise in history, these questions were crucial, in Gunton’s mind, for cultivating a form of life and thought in our modern age, including but not limited to our scientific inquiry. One further preliminary point is necessary by way of introduction to Colin Gunton and his theology. Although attention to the incarnation and to the life, death, resurrection, and ascension of Jesus is the proper See Gunton, Enlightenment and Alienation, 1-2.

4

216

Murray A. Rae

subject matter of theology, theological inquiry necessarily takes us into conversation with all other spheres of human inquiry. The Christian gospel is universal in scope; it makes a claim about the nature and purpose of the universe, about the totality of things and the way they are constituted in relation to God. The task of theology, therefore, has a bearing on all of our human endeavors, including not least our scientific endeavors. This universal scope of the theologian’s interest is evident in Gunton’s claim that “the doctrine of creation is that which provides a common foundation for all the human enterprises we call culture, not just theology but science, politics, ethics and art as well.”5 We will embark shortly on a brief exposition of that doctrine as Gunton develops it, but first let me say a little more about the problematic assumptions about the world and our knowledge of it that were bequeathed to us by the Enlightenment.6 The “Modernist Heresy” and the Displacement of God According to Gunton, “the critical philosophy of the Enlightenment generated a false view of what it is to know, elevating certain modes of human thought to the status of virtual infallibility.”7 This false conception of human knowing was an error, Gunton alleges, which, although it has roots further back in classical Greek philosophy, was initiated in the modern era by thinkers like Francis Bacon and René Descartes, who developed the idea of detached, objective inquiry such as was evident in mathematics. This movement of thought was brought to completion by Immanuel Kant’s fateful “division between pure and practical reason.”8 In consequence, Gunton continues, “all other features of culture tended to be reduced to inferiority.”9 Put simply, it was increasingly assumed that

Colin E. Gunton, A Brief Theology of Revelation (Edinburgh: T&T Clark, 1995), 55. It is important to note here that the era we call the “Enlightenment” cannot be characterized as a single movement of thought or as giving rise to an agreed set of assumptions about the world. “Enlightenment” is a shorthand term by which we designate an era of cultural change in which traditional authorities were called into question and new ones emerged, especially in the wake of substantial advances in scientific inquiry. 7 Gunton, Enlightenment and Alienation, 77. 8 Gunton, Enlightenment and Alienation, 83. 9 Gunton, Enlightenment and Alienation, 78. 5 6

Colin E. Gunton (1941–2003)

217

“science gives knowledge, everything else is relative and uncertain.”10 This was the fate suffered by theology in the wake of the Enlightenment’s celebration of scientific rationality. Theology, it was supposed, could offer no more than the opinions and values of the human subject, whereas science, by contrast, delivered propositions about the world whose truth could be judged independently of any relation to the personal subject.11 Gunton is concerned with two closely connected problems here: an epistemological one and an ontological one. While Enlightenment thinkers were right to reject credulity and superstition, their quest for certainty produced “a view of the human mind that falsified its relation to the world, especially in suggesting that there could be attained an absolute objectivity and impartiality: a God’s-eye-view, so to speak, of reality.”12 This epistemological commitment produces an ontological falsehood—namely, a qualitative division between the knower and the known. The presumption of a God’s-eye view is an assumption of the absolute detachment of the knowing subject from the reality that it purports to know. This misrepresents the relation between the human mind and the world. Descartes’s deliberations on human knowing reveal the problem. His quest for absolute certainty and for indubitable truths drove him to a conception of the knowing subject as an immaterial entity that has no necessary connection to the body—that is, to the material world. Few scientists today would endorse Descartes’s dualism, but the presumption, now beginning to fade, that science delivers a detached, dispassionate, and wholly objective account of what is going on in the world cannot be sustained apart from a dualistic conception of the relation between subject and object. In combating this presumption, Gunton finds an ally in the scientist and philosopher Michael Polanyi. A further problem Gunton identifies with this epistemological misstep is the presumption that the truth about the world can be told in full Colin E. Gunton, The One, the Three and the Many: God, Creation and the Culture of Modernity (Cambridge: Cambridge University Press, 1993), 110. 11 Gunton contends that this supposed distinction between science and theology was exacerbated by the widely influential theology of F. D. E. Schleiermacher. See Colin E. Gunton, “The Truth of Christology,” in Belief in Science and in Christian Life: The Relevance of Michael Polanyi’s Thought for Christian Faith and Life, ed. T. F. Torrance (Edinburgh: Handsel Press, 1980), 91-107, esp. 95. 12 Gunton, Enlightenment and Alienation, 4. 10

218

Murray A. Rae

without assistance from or reference to God. Gunton cites the famous story of Pierre-Simon Laplace (1749–1827) presenting his model of the solar system to Napoleon. When asked by Napoleon why there was no place for God in his model, Laplace allegedly replied, “I have no need of that hypothesis.”13 While there is some debate about the accuracy of this conversation, Laplace’s alleged reply illustrates the deistic view about the relation between God and the world that emerged during the Enlightenment. This notion of a “clockmaker deity” emerged in the wake of an essentially nontrinitarian view of the relation between God and the world that can be traced back to the Middle Ages. As demonstrated in Michael Buckley’s book At the Origins of Modern Atheism, which is frequently cited by Gunton, the nontrinitarian view of God paved the way in turn for modern atheism.14 It helped to create an intellectual climate in which contemporary scientists commonly suppose that the world is wholly explicable without reference to God. There is, however, an awkwardness about such attempts at explanation. As Gunton observes, it is not uncommon to find scientists who are vehemently atheist but who are unable to dispense with the idea of personal agency in their descriptions of nature. In defiance of their own metaphysical commitments, personal agency is commonly attributed to the creation itself. The biologist Richard Dawkins and the chemist Peter Atkins are cited by Gunton as cases in point.15 After quoting evidence from their respective works, Gunton writes, How easily it happens that where God is no longer understood as the overall creator and upholder of the universe there is a reversion to the pagan attribution of agency to the impersonal worlds of molecules, evolution and chaos. The choice is inescapable: either God or the world itself provides the reason why things are as they are. To “personalise” the universe or parts of it, particularly inert substances like molecules, is to succumb to crude forms of superstition.16 Gunton refers to this story in Colin E. Gunton, The Triune Creator: A Historical and Systematic Study (Edinburgh: Edinburgh University Press, 1998), 133. 14 See Michael Buckley, At the Origins of Modern Atheism (New Haven, CT: Yale University Press, 1981). 15 Gunton, Triune Creator, 38. 16 Gunton, Triune Creator, 39. See also 186n14. 13

Colin E. Gunton (1941–2003)

219

Biological determinism is the inevitable result and therewith the elimination of human freedom. Later in the same book, Gunton explains that “if we see the world outside its relation to God, we do not see it properly. . . . When God is not confessed as the Lord of creation, either titanic man, or deified gene take the floor with the result that both understanding and the world are distorted.”17 Gunton has no objection to Darwinism as a theory about the adaptation and evolution of species, but he objects strongly to Darwinism being turned into an all-encompassing dogma about the origin and essential nature of all life. Dawkins claims in the name of science much more than can be justified scientifically, thus providing an easy example of the distortion of science by false theological assumptions. This is not the fault of science as such, however. Science understood in its multiple forms as a mode of inquiry that investigates the causal relations between, and the constitution of, material objects is, in Gunton’s view, a valuable, indeed essential, dialogue partner for theology as it seeks to understand the world. The problem arises—the problem of “scientism”—only when science purports to explain more than it can.18 The kinds of claims made by the likes of Dawkins and Atkins reveal the deeper problem, originating with Galileo, of the world being conceived in mechanistic terms. The essential nature of the world, when conceived as a machine, is thought to be wholly explicable within the confines of material causation. Human freedom, as noted above, is one casualty of this conception of things. Any suggestion of God’s involvement in the world, other than as the power that set the world in motion, is another. The mechanistic view is not entirely without merit; according to Gunton, “it taught the scientist to take the created universe seriously in itself . . . and assisted the emergence of the great classical physics associated with Isaac Newton,” but, “theologically, its long-term effect was disastrous.”19 Of all the claims made in the name of modern science, the denial of divine involvement in the world is the one that Gunton considers to be the most problematic. His trinitarian account of Gunton, Triune Creator, 145. For further comment on this, see Gunton, One, the Three and the Many, 35. 19 Gunton, Triune Creator, 126. 17 18

220

Murray A. Rae

the doctrine of creation, which we will shortly explore, is developed as a counter to this denial of divine agency in the world. This denial was prevalent not only among scientists but also in much of the theological and biblical scholarship being undertaken among Gunton’s contemporaries. It has not yet departed entirely from the realms of biblical and theological scholarship. How then can the Christian doctrine of creation help us to overcome such egregious misrepresentations of reality? The Doctrine of Creation I believe . . . The first thing Gunton wants us to notice about the Christian doctrine of creation is that it is a matter of belief. It is expressed first of all in creedal form: “I believe in God the Father, maker of heaven and earth.” That observation is not a promising beginning for those wedded to an Enlightenment ideal of truth and knowledge. But Gunton is undeterred: the creedal expression of the matter, he says, “shows that the ­doctrine of creation is not something self-evident or the discovery of disinterested reason, but part of the fabric of Christian response to revelation.”20 Put otherwise, the doctrine of creation is not derived from a detached and impartial observation of nature but is the fruit of careful attention to God’s self-disclosure through the Word and the Spirit. The appeal to revelation is of interest not only because God’s self-disclosure is the proper ground of all that theology has to say but also because ­revelation plays a part, in fact, in all our knowing of the world. There is a givenness of things to which we must attend and to which we must allow our minds to be conformed. The language of revelation is appropriate in science as well as in theology. Science proceeds, as its best practitioners attest, by allowing the reality with which it is concerned to be itself, and so reveal itself. To be sure, scientists approach the object of their investigations with a whole range of prior concepts, and they often engage in all kinds of experimental manipulation of those objects in order to test the degree to which the object under investigation conforms to those prior concepts; but true inquiry is undermined if they suppose that the object must be wholly Gunton, Triune Creator, 8.

20

Colin E. Gunton (1941–2003)

221

accommodated within the range of concepts they already have available. Genuine inquiry requires openness to the possibility that their conceptual frameworks may require modification, or replacement, in light of what is shown to them by the object itself. The reality of the object, in other words, precedes and must be allowed, in principle, to subvert our prior classifications of it. In all spheres of human inquiry, we must allow our minds to be conformed to the reality with which we are concerned rather than insisting that the reality under investigation conform to our prior concepts. It is not appropriate in all cases—though it certainly is in the case of persons—to ascribe communicative agency to the object of our inquiries, but the language of revelation is helpful nevertheless in pointing to the givenness of things to which, as theologians and as scientists, we must be attentive and responsible.21 Gunton’s confidence in this account of our cognitive engagement with the world is strengthened by his reading of the scientist and philosopher Michael Polanyi. During the early years of Gunton’s career, Polanyi had become an influential figure, particularly in Great Britain, in discussions about theology and science. The use of ­Polanyi by T. F. Torrance was especially important.22 Torrance found in Polanyi a scientist who was saying the same kinds of things about epistemology that theologians had long held to be true. Polanyi helped to confirm that there was no need for theologians to capitulate to the “modernist heresy” of detached, objective inquiry precisely because that view of the knowing process does not accurately portray either our personal involvement in all of our knowing or the decisive role that antecedent belief plays in our knowing. Polanyi contends that if we are to understand properly what goes on in our knowing of the world, we must recover the teaching of St. Augustine that “all knowledge [is] a gift of grace.”23 He insists further that our quest for knowledge and understanding always On which, see Gunton, Brief Theology of Revelation, 35. See, for example, T. F. Torrance, ed., Belief in Science and in Christian Life: The Relevance of Michael Polanyi’s Thought for Christian Faith and Life (Edinburgh: Handsel Press, 1980). Gunton himself contributed to this collection of essays. 23 Michael Polanyi, Personal Knowledge (London: Routledge/Kegan Paul, 1958), 266. Gunton discusses Polanyi’s appropriation of this principle in Christ and Creation (Carlisle, UK: Paternoster Press, 1992), 14. See also Gunton’s chapter “The Truth of Christology,” in Torrance, Belief in Science and in Christian Life, 91-107. 21

22

222

Murray A. Rae

takes place “under the guidance of antecedent belief.”24 Augustine is again invoked: “Unless you believe, you will not understand.” Like theology, scientific endeavor rests on a series of beliefs about the nature of the universe and its intelligibility; the trustworthiness, though not infallibility, of particular communities of discourse; the sharing of a cultural heritage; and so on. Polanyi explains that “no intelligence, however critical or original, can operate outside such a fiduciary framework.”25 The scientist proceeds within a framework of convictions about the world and about the scientific enterprise—for instance, that the world is ordered and intelligible; that it operates in ways that are relatively stable and predictable; that the internal logic governing its behavior is accessible to us through disciplined inquiry; that our senses provide a means of reliable though not infallible access to material objects; that corroborations of empirical claims usually have their source in laws of nature; that memory can be relied on; and that “the perceptual experiences of other selves are qualitatively similar to or identical with our own.”26 These convictions are not baseless, and they continue to be confirmed and strengthened as scientists go about their work and find them to be cogent, but they are nonetheless beliefs that scientists do not generally think it necessary to prove before going about their work. They simply accept these things as part of the framework of belief within which science operates. Because belief plays a part in all human inquiry, the creedal foundation of the doctrine of creation—“I believe in God the Father, maker of heaven and earth”—is not something to be embarrassed about. It is one further instance, rather, of the essential role of belief of all our knowing of the world. Creatio ex nihilo. The second feature of the Christian doctrine of creation to which Gunton draws attention is the idea of creation out of nothing, creatio ex nihilo. The idea of creatio ex nihilo is not explicitly affirmed in any scriptural text but relies instead on the numerous affirmations in Scripture of divine sovereignty. That sovereignty is absolute Polanyi, Personal Knowledge, 14. Polanyi, Personal Knowledge, 266. 26 Haig Katchadourian, “Some Metaphysical Presuppositions of Science,” Philosophy of Science 22, no. 3 (July 1955): 194-204, 202. 24 25

Colin E. Gunton (1941–2003)

223

and so is not constrained by forces or things that precede God’s creative act or exist independently of it. The absolute sovereignty of God entails that no such forces or things exist. This means that God does not create the world in the manner of a potter, for example, who works with clay to make a pot. For all the creativity, imagination, and freedom that may be exercised by the potter, she is constrained, nevertheless, by the inherent properties of clay. The Christian notion of creatio ex nihilo teaches that the work of divine creation has no such constraints. That is not to say that there are no antecedent conditions that determine the character and form of the created order; but these antecedent conditions are not external to God. Rather, the character and form of creation are determined solely by God’s love, freely exercised in bringing into being that which is other than God.27 This enables us to affirm, following the divine declaration in Genesis 1, that creation is very good. The notion of creation out of nothing enables us also to say, uniquely in the history of human thought, that the world is neither eternal nor infinite. It has a beginning in time and will also have its end. Putting these two affirmations together, we may say both that the world is the outcome of God’s love and that its telos, or true end, lies in the full realization of God’s loving purpose for it. We will say more in due course about what this means for science. But first we need to explore some further entailments of the idea of creatio ex nihilo. To say that the world is solely and completely the outcome of divine love is to say that it is a product of God’s will. The importance of this point cannot be underestimated. To say that the world is a product of divine will means, first, that it need not exist (it is contingent); second, it enables us to say that the world can be other than God (it is not itself divine). This has an important further corollary, as was observed by the sixth-century theologian John Philoponus. Harold Nebelsick explains, Gunton’s stress on the free exercise of God’s love shows the influence of Barth on his theology, although, in one of the last of his books published before he died, Gunton advises that P. T. Forsyth’s stress on the holiness of God’s love is to be preferred to Barth’s stress on freedom, particularly because the holiness of God’s love “encompasses far more adequately the shape of the love as involving the overcoming of the sin that brings men and women into enmity with God.” See Colin E. Gunton, Act and Being: Towards a Theology of the Divine Attributes (London: SCM Press, 2002), 117.

27

224

Murray A. Rae

Following his perception that God was responsible for the creation of the whole universe, Philoponos was convinced that the cosmos as a whole was composed of the same kind of matter and was subject to the same laws. Hence, in direct opposition to prevailing thought, he both rejected the dichotomy between the finite earthly and the infinite eternal heavenly realms and recognised the importance of earthly reality. Further, especially in contrast to the neo-Platonism of his day, Philoponos insisted that nature could not be understood as the finite representation of infinite reality but as real in itself. To understand reality one must make deductions based on observation.28

Philoponus’s Christian insight set him apart from the world of classical thought but not, of course, from Scripture. The biblical story of creation in Genesis 1 insists on precisely the same point—namely, that the heavenly realms, in particular the sun, the moon, and the stars, are not divine beings but are created alongside all others of God’s creatures. It may be significant too for the biblical author, in order to avoid any suggestion of a hierarchy of being, that the celestial bodies do not appear until the fourth day. The heavens are made of the same stuff as earthly realities, and none of them are divine. This opens the way to scientific (that is to say, empirical) investigation of what created realities consist in and how they behave. Gunton thus explains that “without the doctrine of creation out of nothing which affirmed the rationality, contingency and non-divinity of the material world, the rational and experimental techniques which have brought such immense enrichment of human culture simply would not have been.”29 We should note in passing that scientific inquiry in the modern sense is a very different kind of enterprise than the “science” said to have been undertaken in antiquity. While in both ages there are those who seek understanding of “how the heavens go” to invoke the phrase appealed to Harold Nebelsick, The Renaissance, the Reformation and the Rise of Science (Edinburgh: T&T Clark, 1992), 113, italics original. Cited in Gunton, Triune Creator, 72-73; cf. Gunton, Brief Theology of Revelation, 53-54. 29 Gunton, Brief Theology of Revelation, 57. Gunton finds support for this view in Michael Foster’s 1934 article, “The Christian Doctrine of Creation and the Rise of Modern Natural Science” Mind 43, no. 172 (October 1934): 446-68. See especially the several references to this article in Gunton, One, the Three and the Many. 28

Colin E. Gunton (1941–2003)

225

by Galileo,30 Peter Harrison has recently shown that “the ancient Greeks had neither activities nor occupations that are directly equivalent to our terms ‘science’ and ‘scientist.’”31 The natural philosophy commonly regarded as the forerunner of modern science was a branch of philosophy in general that was concerned, above all, with the art of living. Harrison explains that “a connection between the study of nature and the philosophical life was provided by the assumption that a moral order is built into the structure of the cosmos.”32 Consider, for example, Plato’s advice in the Timaeus: And the motions which are naturally akin to the divine principle within us are the thoughts and revolutions of the universe. These each man should follow, and by learning the harmonies and revolutions of the universe, should correct the courses of the head which were corrupted at our birth, and should assimilate the thinking being to the thought, renewing his original nature, so that having assimilated them he may attain to that best life which the gods have set before mankind, both for the present and the future.33

Not all of this is quite as a Christian theologian might put it, but there are two points of interest for our purposes; the first is Plato’s conviction that there is in the universe a divinely bestowed order and intelligibility, and the second is the connection between scientific inquiry, to use our modern term, and moral or personal formation. That is a point to which I shall return. In the meantime, let us return to the substantive point that certain things need to be believed about the universe in order for science to get In his letter to the Grand Duchess Christina of Tuscany in 1615, Galileo wrote, “I would say here something that was heard from an ecclesiastic of the most eminent degree: ‘That the intention of the Holy Ghost is to teach us how one goes to heaven. not how heaven goes.’” Galileo Galilei, “Letter to the Grand Duchess Christina of Tuscany, 1615,” Internet History Sourcebooks Project, August 1997, https://sourcebooks.fordham.edu/halsall/mod/galileo-tuscany.asp. 31 Peter Harrison, The Territories of Science and Religion (Chicago: University of Chicago Press, 2015), 26. Harrison points the reader to Rémi Brague, The Wisdom of the World: The Human Experience of the Universe in Western Thought, trans. Teresa Fagan (Chicago: University of Chicago Press, 2003). 32 Harrison, Territories of Science and Religion, 27. 33 Plato, Timaeus 90d, from The Collected Dialogues of Plato, ed. Edith Hamilton and Huntington Cairns (Princeton, NJ: Princeton University Press, 2005), cited by Harrison, Territories of Science and Religion, 28. 30

226

Murray A. Rae

under way. Gunton puts the matter thus: “Science could not come to be until it came to be believed that the structures of material reality, the world presented to the mind through the senses, were intelligible in their contingent relations.”34 The universe must be understood as relatively independent of God—it must have its own distinct being, and it must be regarded as having an order and intelligibility that is, in principle at least, accessible to human inquiry. All of these conditions are supplied, Gunton explains, by the Christian doctrine of creation. Whether modern science could have arisen elsewhere or taken the particular form it has in cultures not shaped by the Christian doctrine of creation, I leave to one side;35 the salient point is that scientific endeavor is not in itself antithetical to a theological construal of the world, but it depends, in fact, on beliefs that Christian theology has long held and helps to make plausible. This point is completely independent of what individual scientists may believe or deny about God’s involvement in the creation and ordering of the universe. As it happens, Gunton was fond of citing Michael Faraday in support of the view that the laws of physics and chemistry were willed by God at the creation,36 but there is no shortage of scientists who deny the point. Popular vote cannot settle the matter; moreover, insofar as science concerns itself with the empirical observation of natural phenomena, the question of whether God is involved in establishing and upholding the laws of nature is not a matter that science can settle.37 One more point must be treated within this discussion of creatio ex nihilo. It is a point already mentioned but which we now retrieve for fuller consideration—namely, that the world has a beginning and also an end. To confess that the world is the product of God’s will and that its form and character are determined by divine love is to say that it has Gunton, Brief Theology of Revelation, 57. Although Gunton stresses the important role played by Christian theology, he does acknowledge that “we must beware of claiming along with some modern theologians too much of the credit for the development of modern science.” Colin E. Gunton, “Relation and Relativity,” in The Promise of Trinitarian Theology, 2nd ed. (Edinburgh: T&T Clark, 1997), 137-57, 155. 36 See, for example, Gunton, Brief Theology of Revelation, 59, although here he cites Geoffrey Cantor’s account of Faraday rather than Faraday himself. 37 For the same reason that theological judgments are not amenable to empirical proofs, Gunton rejects natural theology. General revelation and a theology of nature are a different matter; see Gunton, Brief Theology of Revelation, 60-63. 34 35

Colin E. Gunton (1941–2003)

227

some purpose. God has created with an end in view. Gunton repeatedly stresses that eschatology is intrinsic to the doctrine of creation and not a separate field of theological deliberation. There is, he says, no creation “in the beginning” without an eschatological orientation. From the beginning, it has a destiny, a purpose. Creation is “out of nothing” in that it is made both to be and to become something, not something else but something perfected, able to praise and give glory to God for what it has been enabled by him to become.38

Simply put, creation is a project. It is a project initiated by God that will in time be brought to fulfilment by God, but it is a project established for the sake of the creature in which the creature is called to participate. Because the form and character of creation is determined by divine love, our participation is not coerced but enabled, not unlike the way a child may grasp a parent’s extended hand as they embark on some project together. The child’s act is freely undertaken but made possible in virtue of the parent’s love. In the case of creation, the telos to which creation is directed is the flourishing and harmonious coexistence of all that God has made in loving communion with its Creator. Creation is, on this account, the space and time given for the creature to grow up in love for its Creator. Gunton here appeals to the secondcentury theologian Irenaeus, who, as we will see shortly, is the theologian who most profoundly influenced Gunton’s account of the doctrine of creation. Irenaeus’s eschatology is distinctive on account of his insistence that, although good, Adam is not created perfect.39 That is why, for Irenaeus, the final realization of God’s creative purposes cannot be conceived as a return to some ideal state imagined to have existed in the beginning. As Gunton puts it, “For Irenaeus, that which comes from nothing is destined to become something.”40 The orthodox tradition speaks of our place in creation as the priests of creation, gathering all things into communion with their Creator. Another part of the Christian story tells of how we have failed in this responsibility but that it has been taken up and exercised for Colin E. Gunton, The Christian Faith: An Introduction to Christian Doctrine (Oxford: Blackwell, 2002), 19. 39 See Gunton, Triune Creator, 55-56. 40 Gunton, Triune Creator, 55, italics original. 38

228

Murray A. Rae

us by Christ, and in a way that restores the possibility of our sharing through the Spirit in this divine calling and appointment. The connection in Irenaeus between the divine project of creation and the tragedy of human sin is explained by Douglas Farrow: The love for God which is the life of man cannot emerge ex nihilo in full bloom; it requires to grow with experience. But that in turn is what makes the fall, however unsurprising, such a devastating affair. In the fall man is “turned backwards.” He does not grow up in the love of God as he is intended to. The course of his time, his so-called progress, is set in the wrong direction.41

Commenting on this passage, Gunton explains that “redemption or salvation is that divine action which returns the creation to its proper direction, its orientation to its eschatological destiny, which is to be perfected in due course of time by God’s enabling it to be that which it was created to be.”42 This leads us to the third and most distinctive claim of the Christian doctrine of creation—namely, that creation is the work of the triune God: Father, Son, and Holy Spirit. Creation and the triune God. I have pointed out already the importance for science of upholding the relative independence and ontological distinction of the creation from the Creator. The world is not in any sense divine but has its own distinct existence. Because God does not need us, the creature has freedom and value in its own right. That God is wholly other raises the question, however, of how God and the world are related. By what means does God bring the world into being and continue his relationship with it? The progress and success of science in the early modern period persuaded some thinkers of the time that the world’s operation according to natural laws left no room for divine involvement in creation. Deism, as already noted, holds that God was responsible for bringing the world into being and establishing the laws by which it operates, but the laws of nature now function without the need of further divine involvement. According the English deist Thomas Burnet, “We think him a better Artist that makes a Clock that strikes regularly at Douglas Farrow, “St Irenaeus of Lyons: The Church and the World,” Pro Ecclesia 4, no. 3 (August 1995): 333-55, 348. Cited in Gunton, Triune Creator, 56. 42 Gunton, Triune Creator, 56. 41

Colin E. Gunton (1941–2003)

229

every hour from the Springs and Wheels which he puts into the work, than he that hath so made his Clock that he must put his finger to it every hour to make it strike.”43 While the deist view, as a formal philosophical position, never gained widespread support, the assumption that science has displaced God and rendered him redundant is commonplace in contemporary Western culture. There is work to do in the conversation between science and theology on how we may speak of divine agency in ways that both respect the causal regularities that science investigates yet uphold the biblical conviction that God is at work sustaining all things by his power and directing them toward their proper end in reconciled communion with him. Gunton’s contribution to this conversation lies especially in his insistence that the Creator, Sustainer, and Perfecter of all things is the triune God—Father, Son, and Spirit. How does this help? In the work of Irenaeus, Gunton finds the conceptual resources to speak of God’s relation to the world. Irenaeus’s theology, forged in opposition to those heresies grouped under the name of Gnosticism,44 is an exercise in thinking through what must be said of the world and of God in light of the incarnation and of the life, death, and resurrection of Jesus. What must be said first of all, in opposition to the Gnostics, is that the material world is good. “If God in his Son takes to himself the reality of human flesh,” says Gunton, “then nothing created, and certainly nothing material, can be downgraded to unreality, semi-reality, or treated as fundamentally evil, as in the Manichaean version of Gnosticism.”45 The incarnation demonstrates, furthermore, along with the ensuing life, death, and resurrection of Jesus, that this world that we inhabit is the terrain within which God brings about his purposes. The relation between God and world is therefore to be thought through Christologically (in the light of Christ) and, as we shall also see, pneumatologically (with reference to the Spirit). It is in the light of Christ also that Irenaeus affirms that God creates out of nothing. If it were not so, God would be constrained in his creative Thomas Burnet, The Sacred Theory of the Earth, 3rd ed. (London: Walter Kettilby, 1697), 72. I take the point from Gunton, Triune Creator, 52. 45 Gunton, Triune Creator, 52. 43 44

230

Murray A. Rae

work; he would be, in Irenaeus’s words, a “slave to necessity,”46 and his omnipotence would be compromised. The notion of omnipotence sometimes raises concerns for those wishing to uphold a scientific view of the universe, for it suggests, in the minds of some, a capacity to interfere arbitrarily in the workings of the created order. But that is not at all how we should think of omnipotence. We should instead understand the term in the light of Christ. For Irenaeus, the omnipotence of God signifies that God does not “stand in the need of other instruments for the creation of those things which are summoned into existence. His own Word is both suitable and sufficient for the formation of all things.”47 This reference to creation through his Word gives a first hint of the theology of mediation that Irenaeus develops and which Gunton himself considers to be crucial to understanding God’s relation to the world. “As is well known,” Gunton writes, “Irenaeus frequently says that God creates by means of his two hands, the Son and the Spirit. This enables him to give a clear account of how God relates to that which is not God: of how the creator interacts with his creation.”48 Two points are of particular interest here: first, God’s involvement with the world is personal involvement. God needs no intermediaries. His creation of the world, along with his involvement in the world’s redemption and final reconciliation, is undertaken in person through his own Word and Spirit. Irenaeus says, It was not angels . . . who made us, nor who formed us . . . nor any Power remotely distant from the Father of all things. For God did not stand in need of these [beings], in order to the accomplishing of what he had himself determined with Himself beforehand should be done, as if He did not possess his own hands. For with Him were always present the Word and the Wisdom, the Son and the Sprit, by whom and in whom freely and spontaneously, He made all things.49 Irenaeus, Against Heresies 2.5.4, in Alexander Roberts and James Donaldson, eds., Ante-Nicene Fathers, rev. ed. (Peabody, MA: Hendrickson, 1994), 1:365. I owe the reference to Gunton, Triune Creator, 53. Subsequent references to Against Heresies will show the book, chapter, and paragraph number, followed by the page number in the Hendrickson edition. 47 Irenaeus, Against Heresies 2.2.5, 361. 48 Gunton, Triune Creator, 54. 49 Irenaeus, Against Heresies 4.20.1, 487-88.

46

Colin E. Gunton (1941–2003)

231

The second point deriving from Irenaeus’s concept of mediation, as revealed through the incarnation, is that God works within and according to the conditions of created existence rather than against them. The Word becomes flesh and dwells among us (Jn 1:14) and so takes on the full reality of creaturely existence. God’s involvement with the world does not take the form of arbitrary disruptions of created reality or violations of its order. While creation is spoken of by Gunton as the work of the triune God, and while in all things the Father, Son, and Spirit work together, it is the Spirit in particular who is described in Scripture and in the Nicene Creed as the giver of life.50 The Spirit is the one who animates creation and breathes life into the creature. Echoing Genesis 2:7, which speaks of God forming man from the dust of the ground and breathing into his nostrils the breath of life, the psalmist declares, “When you send your Spirit [ruach], they [all living things] are created” (Ps 104:30) and “when you take away their breath [ruach], they die” (Ps 104:29). Whatever the details of the evolutionary mechanism by which life has come to be on this earth—about which science can teach us a great deal—life is affirmed by the psalmist as given and sustained by God. It is not mere biological existence that the Spirit gives and sustains, however, but life in all its fullness. Taking his lead at this point from the fourth-century theologian Basil of Caesarea, Gunton speaks of the Spirit as the perfecter of creation.51 Paul’s talk of the gifts and fruits of the Spirit also bears on this topic because it speaks of the Spirit enabling human beings to become what they are intended to be by God. The fruits of the Spirit—love, joy, peace, patience, kindness, and so on (Gal 5:22), along with the gifts of wisdom, knowledge, faith, and healing (1 Cor 12:7-11)—are gifts enabling the full realization of our humanity. This work of the Spirit is not therefore a violation of our freedom, nor of the laws of nature; rather, it is the operation of God within the conditions of created existence realizing his promise that the creature shall have life in all its fullness.

So the Nicene Creed: “I believe in the Holy Spirit, the Lord, the giver of life.” Cf. Job 33:4; Ezek 37:10; Rom 8:2, 11; 2 Cor 3:6. 51 See, e.g., Gunton, Triune Creator, 10, 86. The reference is to Basil, On the Holy Spirit 15.38. 50

232

Murray A. Rae

Science and Theology in Collaboration More can be said about the triune mediation of creation and of the means through which God involves himself in our world, but we turn now to develop further some of the indications given along the way of what all this means for the way we might think about science. In this last section of the essay I wish to explore briefly four points concerning the relation between science and theology that emerge from Gunton’s consideration of the doctrine of creation. The conditions for scientific inquiry. The first point has largely been covered already but is worth repeating, especially because it is not commonly understood. The Christian doctrine of creation, as Gunton has shown, expresses in theological mode the conceptual conditions under which scientific endeavor may get under way. Chief among these is the divinely established order and intelligibility of the universe and the accessibility of that order to human inquiry. That includes the empirical and rational enquiries of science, but not exclusively so.52 The order and intelligibility of the universe is accessible also through artistic endeavor and in varying degrees through our daily experience as creatures given the responsibility of stewardship of the creation. That divine calling and appointment entails that we have the capacity, also God-given and to be developed over time, to understand the workings of the world. This is part of what is affirmed in Genesis 2:19-20 where humanity is tasked with naming the other living creatures. The endeavors of science are further supported by the biblical insight that neither the world nor any part of it is divine. It has its own particular being and integrity and so can be investigated on its own terms. We saw the importance here of early Christian thinkers like John Philoponos, who insisted, in contrast with the surrounding culture of antiquity, that the heavenly bodies were made of the same stuff as earthly bodies and were subject to the same kind of laws. While progress in science relies heavily on empirical investigation of the world, substantial progress in science can also be made by other means. A. N. Hunt has recently explored the testimony of scientists who claim that they were persuaded of the truth of particular scientific theories not by empirical proofs but by an appreciation of the beauty of the theory. See A. N. Hunt, “Ciphers of Transcendence: Cognitive Aesthetics in Science,” The Heythrop Journal 49 (2008): 603-19. Consider also Thomas Dubay’s investigation of the “evidential power” of beauty. See Thomas Dubay, The Evidential Power of Beauty: Science and Theology Meet (San Francisco, CA: Ignatius Press, 1999).

52

Colin E. Gunton (1941–2003)

233

The fruitfulness of this kind of Christian thinking for scientific inquiry is often denied. We have already mentioned Peter Harrison’s debunking of the widely believed myth that the Christianity of the Middle Ages posed a major obstacle to scientific progress that was revived only when the scientific revolution of the seventeenth century enabled science to break free of religious shackles.53 We may cite as well Stephen Gaukroger’s well-supported claim that “Christianity took over natural philosophy in the seventeenth century, setting its agenda and projecting it forward in a way quite different from that of any other scientific culture.”54 Isolated incidents of tension between scientific and ecclesial authority have unfortunately obscured the very substantial contribution that theological claims about the nature of reality have made to the progress of science. A corollary of the claim that the Christian doctrine of creation includes affirmations about the world that are germane to the development of scientific inquiry is that we should not be surprised to find a degree of complementarity between theological and scientific claims about the nature of reality. Gunton is fond of citing particular examples. Reflecting on the perichoretic relations of the three persons of the Trinity, for instance, and the lessons that may be learned about the relational and ­interdependent character of all things, Gunton observes a degree of complementarity between this theological observation and similar claims made by much modern physics and cosmology. “Many scientists,” he observes, “have spoken the language of perichoresis in their descriptions of the universe.”55 Scientist Michael Faraday and Ilya Prigogine, a physical chemist and Nobel laureate, are cited in support of the point, the latter in concert with philosopher of science Isabelle Stengers. Prigogine and Stengers write that “[Physics] now recognizes that, for an interaction to be real, the ‘nature’ of the related things must derive from these relations, while at the same time the relations must derive from the ‘nature’ See again Peter Harrison, The Territories of Science and Religion. Herbert Butterfield is cited as a leading proponent of this myth (See Herbert Butterfield, The Origins of Modern Science, 1300– 1800, 2nd ed. [New York: Macmillan, 1962]), which is also promoted with much enthusiasm and no little ignorance of history by such as Richard Dawkins and Christopher Hitchens. 54 Stephen Gaukroger, The Emergence of a Scientific Culture: Science and the Shaping of Modernity, 1680–1760 (Oxford: Oxford University Press, 2010), 23. Cited in Harrison, Territories of Science and Religion, 189. 55 Gunton, One, the Three and the Many, 172. 53

234

Murray A. Rae

of the things.” That, for Gunton, “is a statement of created, analogous, perichoresis. Everything in the universe is what it is by virtue of its relatedness to everything else.”56 In the light of the doctrine of creation elucidated above, such convergences between theological and scientific claims about reality should come as no surprise. The use of Faraday and of Prigogine here is typical of the kinds of engagement with science that we find scattered throughout Gunton’s writings. While being very clear about the errors of scientism, he is genuinely interested in scientific insight into the nature of reality and appreciates the congruence he commonly finds between theological and scientific explanation of what is going on in the world. Although Gunton more explicitly and extensively engages issues related to the sciences, it is also worth noting, if only in passing, that Gunton expects the same kind of congruence in the arts. All three fields of human inquiry and representation are capable of reflecting the deep truth of things and should expect therefore to discover points of convergence in their respective endeavors.57 Theological and scientific explanation. We have seen the appeal Gunton makes to the Irenaean metaphor that God creates sustains, redeems, and brings the world to perfection through his two hands, the Son and the Spirit. This is a theological claim and a statement of belief rather than empirical investigation. What relation does such a statement have to the kinds of statements about the world that emerge from the sciences—that the universe got under way with the “Big Bang,” for instance, or that biological life has evolved through the natural selection and gradual mutation of species? The first thing to note is that theology and science,58 though both concerned with the nature of reality, are not rivals, although they can be and often have been presented as such. The disciplines of theology and science involve two different kinds of judgments about how reality is constituted. This claim and the preceding citation from Prigogine and Stengers appears in Gunton, One, the Three and the Many, 172. The citation is from Ilya Prigogine and Isabelle Stengers, Order out of Chaos: Man’s New Dialogue with Nature (London: Fontana, 1985), 95. 57 See, Gunton, One, the Three and the Many, 175. 58 I am using the term science here in a very generic sense to indicate the empirical investigation of created realities. 56

Colin E. Gunton (1941–2003)

235

Both science and theology are captivated by the order of things, and both venture an explanation. Theological explanations—Genesis 1 is exemplary here—testify that because God is the author and sustainer of the world’s order, and because that order is the outcome of divine love, that order can be relied on to serve God’s loving purposes for the creature. Science, for its part, investigates the material means by which that order has developed and through which it is sustained, and so offers, through observation of the natural world, its own complementary assurance that the order of things can be relied on. So confident has science become of the reliability of the world’s order that it happily speaks of the laws of nature. This assurance of reliability and consistency in the workings of the universe makes possible a vast range of human projects, some of them consonant with but others antithetical to God’s good purposes for the world. Theology, for its part, proposes, as we have seen, that the very logic and order of the universe, its reason for being and the purpose to which it is directed, are revealed, most especially, in Christ. To put it in the language of the apostle Paul, Christ is the one through and for whom all things came to be and in whom all things hold together (See Col 1:15-20). It is in John’s Gospel, however, that we see the most profound expression of this theological conviction. The Gospel begins, “In the beginning was the Word [the Logos], and the Word was with God, and the Word was God. He was with God in the beginning. Through him all things were made; without him nothing was made that has been made” (Jn 1:1-3). Then in verse 14, an extraordinary claim is made: this Logos, this Word, “became flesh and made his dwelling among us.” All the order that science discerns, the constancy and the reliability of the workings of nature, has its basis in the Christ we encounter in Jesus of Nazareth; it is through Jesus that we may discern to what purpose creation’s order is directed. As John’s account of Jesus’ life unfolds, we learn that the order of creation is directed to a particular end; that end is fullness of life for the creature in communion with God. As John 1:3-4 tells us, what has come into being in Christ is life. Again, in John 10:10, he has come that we might have life and have it abundantly. The wondrous workings of the universe that science so impressively explores are directed to this end, that the creature

236

Murray A. Rae

may have life in all its fullness. The gospel tells of the various ways in which humanity has chosen death rather than life and of the disruptions to the created order brought about as a result. That story takes longer to tell than we have time for here, but let us note that in John’s Gospel especially, Jesus’ ministry involves the restoration of the creation to its proper purpose. His turning of water into wine at the wedding in Cana, for example (Jn 2:1-11), is a restoration of divine blessing, reversing the tragedy spoken of in Isaiah 24:7, when the earth was barren, the vines had withered, and there was an outcry in the streets for lack of wine. The healing of a man born blind in John 5 is the overcoming of those disruptions of creaturely life that are manifest in sickness and disease. The feeding of the five thousand in John 6 demonstrates the sufficiency of the earth’s fruitfulness so that, as God intended, there is food enough for all. And so on. John offers us these “signs” that God is at work in Jesus, setting the creation again on its trajectory to fullness of life in Christ. We can trace as well through John’s Gospel the workings of the Spirit, who is the giver of life. All of this comes to its climax in the death and resurrection of Christ. Death represents the severest disruption of God’s good purposes for the creature, but in the Spirit’s raising of Jesus from the grave, even death is overcome. None of this is a rival to scientific explanations of the world. The theology of John’s Gospel speaks rather of the underlying logic and purpose of the workings of the world that science investigates and so provides the framework within which all human endeavors—science, culture, art, and industry—may be properly understood. God’s involvement in the created order—a noncompetitive account. God’s involvement in the world through Word and Spirit enables us to develop a noncompetitive account of divine and creaturely agency. An objection to divine involvement in the workings of the world, made famous especially by the philosopher David Hume, is that God’s involvement in the world, especially through what we call miracles, necessarily involves a violation of the intrinsic order and integrity of the natural world, or, we might say, of the world that science investigates. Hume defined a miracle as a violation of the laws of nature and argued that such violations are so improbable as to be safely discounted. The argument is

Colin E. Gunton (1941–2003)

237

extended by some to suggest that science has shown that miracles cannot happen, the assumption being that we can have a regular and ordered universe or we can have divine action in the world, but we cannot have both. There are several problems with this line of argument, not least among which are the problematic definition of miracle as a violation of the laws of nature and the conception of God as a remote monad who might occasionally interfere with the workings of the world.59 The argument against miracles in the Humean tradition pays no attention, moreover, to how God in fact exercises his agency through the Son and Spirit. Gunton’s doctrine of creation has helped us to see that the order of creation, its inner logic and reason for being, is created and sustained and will ultimately be realized in full by God. That logic, or Logos, in and through which all things have come to be, has become incarnate in Jesus. In that light we can see Jesus’ healing miracles, for instance, not as a violation of the laws of nature but as the restoration of God’s intended order. The Creator intends that the lame shall walk, the blind see, and the hungry be fed. Far from being a violation of the intrinsic order of creation, these acts of God are a restoration of the divine ordering of things; they are the gift of new creation setting right the disruptions to creation’s order that have scarred our fallen world. Likewise, the Spirit poured out at Pentecost does not disrupt but rather renews creaturely life, thus enabling humanity particularly in this case to be “redirected to its proper end,”60 to become what it is intended to be. According to the Christian doctrine of creation, the so-called laws of nature that science helps us to understand and respect are established for the sustenance of creaturely life. They are not violated but are in fact upheld when through the work of Word and Spirit God ensures that his promise of life for the creature is fulfilled. The noncompetitive way in which God acts in the world may be further demonstrated by a couple of biblical examples. An important but often overlooked feature of the account of creation given in Genesis 1 is the permissive nature of God’s creative Word. We strike it first in verse John Earman, Hume’s Abject Failure: The Argument Against Miracles (Oxford: Oxford University Press, 2000). 60 Gunton, Christian Faith, 7. 59

238

Murray A. Rae

3: “Let there be light.” But notice especially its import in verse 11, “Let the land produce vegetation: seed-bearing plants and trees on the land that bear fruit with seed in it, according to their various kinds,” and in verse 20: “Let the water teem with living creatures, and let birds fly above the earth.” Gunton points out that “parts of the earth are empowered to serve as mediators of God’s creation of other parts. . . . Worldly agencies are enabled by divine action to achieve their own ‘subcreating,’ not in the absolute way that God creates, but relatively, as creation from what already is.”61 The earth and the sea are summoned to bring forth life, by natural processes, we might say, but not without God’s enabling. Again, this account of things does not speak of the violation of nature’s “laws” but rather of creation’s divinely established capacity to share in the generation of life. In this way, it is appropriate to speak of divine action and natural law operating together in the evolution of creaturely life.62 A second example of noncompetitive divine and creaturely agency occurs in the conception and birth of Jesus. Mary responds to the angelic annunciation that she will conceive and bear a Son with the words, “Let it be to me according to your word” (Luke 1:38 NKJV). Mary’s response is the proper response of all creaturely life. It is the tragedy of human sin that we so often seek to foist on the world a word and an order of our own making, sometimes with catastrophic results. Human sinfulness rather than divine action is the true threat to creation’s goodness and order. Returning to the biblical account of Jesus’ conception, it is the life-giving Spirit who fashions the child in Mary’s womb (Mt 1:18; Lk 1:35). Gunton comments, “In shaping from the clay of earth a body for the Son, the Spirit enables this part of earth to be fully itself, to move to perfection rather than to dissolution. The point of the doctrine of the virgin birth,” Gunton adds, “is . . . to link together divine initiative and true humanity. Jesus is within the world as human, and yet as a new act of creation by God.”63 The Spirit in this episode, as elsewhere, works with rather than against the conditions of creaturely existence. Gunton, Christian Faith, 7-8. It should be obvious that this biblical account of things happily accommodates the theory of evolution. 63 Gunton, Christ and Creation, 52-53. 61 62

Colin E. Gunton (1941–2003)

239

These biblical and theological claims are not empirical claims of the kind that science can confirm or deny, but neither do they represent any threat to the scientific view of the world. They indicate, rather, that the fine-tuned order of the universe, about which science tells us so much, is the framework within which God is at work bringing his purposes for creation to completion. The context and telos of scientific endeavor. The Christian doctrine of creation involves, as we have seen, strong claims about the end or telos to which creation is directed. That telos—the fullness of creaturely life in loving communion with God—is revealed most especially in Jesus, who with the Father and the Spirit works to overcome the threats to that telos posed by human fallenness and by creation’s bondage to decay. Theologically understood, this history of creation, redemption, and final fulfillment is the context in which all creaturely life and all human endeavor takes place, however much we fail to recognize or to acknowledge it.64 That it is so nevertheless requires that we consider the degree to which our creaturely endeavors share in or frustrate the working out of God’s purposes for the world. There are ethical questions to be considered here. We must ask, for instance, about the degree to which our science serves the well-being of all creation, about the degree to which it is used for healing and for ensuring that the divine provision for the flourishing of creaturely life is accessible to all. We must ask whether the alterations we make to the fabric of creation conform to or disrupt God’s good ordering of things; we must consider whether our scientific and technological endeavors enable us to bequeath the world to those who come after us with its character as blessing still intact, and so on. Our best efforts in science and in theology should therefore be dedicated to understanding the workings of God’s creation. Just as in Paul’s vision of the church, in which all the gifts are needed and all parts of the body have a role to play, wisdom in such questions as I have just outlined requires the combined insight of all scholarly disciplines and all modes of human understanding. It also requires sustained and prayerful attention to the Word and Spirit of God. The tragedy of human sin is that for much of the time given us to share in the reality of God’s creative and life-giving purpose for the world, we go about in ignorance, indifference to, or willful defiance of it.

64

240

Murray A. Rae

I have tried to show that the Christian doctrine of creation reveals the possibility, and indeed the necessity, of respectful collaboration between science and theology. I have also sought to point out the ways in which Gunton himself engaged with science. He engaged sometimes as a critic of what he referred to as “scientism,” the illegitimate presumption that the methods of science are sufficient to comprehend the whole of what can be known. More positively, while Gunton did not actually engage in any direct collaboration with scientists, he drew frequently on their insights wherever he saw a constructive complementarity with theological claims. Such complementarity was not at all surprising for Gunton. Indeed, it is precisely what should be expected in the light of a properly formulated (that is to say, trinitarian) doctrine of creation. Recommended Reading Primary Works Gunton, Colin E. A Brief Theology of Revelation. Edinburgh: T&T Clark, 1995. ———. Christ and Creation. Grand Rapids, MI: Eerdmans, 1992. ———. “The Doctrine of Creation.” In The Cambridge Companion to Christian Doctrine, edited by Colin E. Gunton, 141-57. New York: Cambridge University Press, 1997. ———. Enlightenment and Alienation: An Essay Towards a Trinitarian Theology. Grand Rapids, MI: Eerdmans, 1985. ———. The One, the Three and the Many: God, Creation and the Culture of Modernity. New York: Cambridge University Press, 1993. ———. The Triune Creator: A Historical and Systematic Study. Grand Rapids, MI: ­Eerdmans, 1998.

Secondary Works Harvey, Lincoln. ed. The Theology of Colin Gunton. New York: T&T Clark, 2012. McNall, Joshua. A Free Corrector: Colin Gunton and the Legacy of Augustine. Minneapolis: Fortress Press, 2015. Whitney, William B. Problem and Promise in Colin E. Gunton’s Doctrine of Creation. Vol. 26. Studies in Reformed Theology. Boston: Brill, 2013.

A F T E RWO R D Alister E . McGrath

M

any reasons can be given for encouraging a dialogue between the natural sciences and Christian theology—such as a sense of intellectual curiosity, frustration over the arbitrariness of disciplinary divisions, or a desire for a deeper vision of our world than is offered by theology or the natural sciences on their own. The philosopher Mary Midgley is one of many to argue that our universe is too complex and rich to be captured fully by any one tradition of inquiry.1 We need to make use of many “intellectual toolkits” in our encounter with reality, realizing the strengths and limits of each. We should use multiple angles of approach, offering us different windows into our world, and then try to integrate these multiple insights into a grander vision of our world. This welcome and significant collection of studies is an important resource for both scientists and theologians who rightly sense that there is much to be gained from a dialogue between these two disciplines. According to the historian of science Thomas Dixon, the myth of the “warfare” of science and religion was a self-serving myth that was invented by Enlightenment rationalists in the late 1700s, propagated by Victorian free-thinkers in the late 1800s, and is defended today by “scientific” atheists and other voices competing for authority within Western popular culture.2 A growing awareness that this once-influential “warfare” or “conflict” model of the relation of science and religion was shaped and motivated by cultural agendas and vested interests of the late nineteenth See especially Mary Midgley, Science as Salvation: A Modern Myth and Its Meaning (London: Routledge, 1992). For a full analysis, see Alister E. McGrath, “The Owl of Minerva: Reflections on the Theological Significance of Mary Midgley,” Heythrop Journal 61, no. 5 (2020): 852-64. 2 Thomas Dixon, Science and Religion: A Very Short Introduction (Oxford: Oxford University Press, 2008, 9). 1

242

Alister E. McGrath

century—and hence is an outdated social construction—has led to a new interest in exploring the interface between science and faith, especially on the part of Christian natural scientists who want to hold together their personal faith and their professional calling.3 Although the rise of the New Atheism associated with Richard Dawkins and others briefly rekindled media interest in the “warfare” narrative from about 2006 to 2009, this movement has now been widely discredited, with many former adherents renouncing its superficial and simplistic rhetoric.4 This welcome development can be expected to lead to a wider interest in alternative ways of considering the relationship of science and faith within churches and wider culture. Yet we need more than a changed cultural mood to bring about a deepened engagement between the natural sciences and Christian theology. One of the most pressing needs is to help natural scientists immerse themselves in the riches of Christian theology. Perhaps my own story might illustrate this problem and help explain the importance of this collection of essays.5 As a teenager, focusing on the natural sciences at high school, I found myself drawn to the “warfare” narrative. Such was the imaginative power of this narrative that I concluded that to be a scientist required me to be an atheist. Happily, my atheist phase only lasted a few years, as I slowly began to realize its intellectual vulnerability and existential bleakness. I discovered Christianity in my first year at Oxford University in late 1971, studying chemistry. For some representative assessments, see Jeff Hardin, Ronald L. Numbers, and Ronald A. Binzley, eds., The Warfare Between Science and Religion: The Idea That Wouldn’t Die (Baltimore: Johns Hopkins University Press, 2018); Peter Harrison, “‘Science’ and ‘Religion’: Constructing the Boundaries.” Journal of Religion 86, no. 1 (2006): 81-106; Bernard Lightman, ed., Rethinking History, Science, and Religion: An Exploration of Conflict and the Complexity Principle (Pittsburgh: University of Pittsburgh Press, 2019); Joshua M. Moritz, “The War That Never Was: Exploding the Myth of the Historical Conflict Between Christianity and Science,” Theology and Science 10, no. 2 (2012): 113–23; Alvin Plantinga, Where the Conflict Really Lies: Science, Religion, and Naturalism (New York: Oxford University Press, 2011). 4 An excellent example is P. Z. Myers, who formerly identified himself as a New Atheist, but later disavowed this movement: see his January 2019 blog post, “The Train Wreck That Was the New Atheism,” https://freethoughtblogs.com/pharyngula/2019/01/25/the-train-wreck-that-was-the -new-atheism. 5 I tell the story of my transition from an atheist scientist to a Christian theologian in a short memoir: Alister McGrath, Through a Glass Darkly: Journeys Through Science, Faith and Doubt— A Memoir (London: Hodder & Stoughton, 2020). 3

Afterword

243

Initially, I tried to develop or hoped to discover intellectual frameworks that might help me hold together my developing Christian faith and my continuing love of science. In the end, I found such a framework through hearing Charles A. Coulson preaching on the theme of science and faith. Coulson was then Oxford’s Professor of Theoretical Chemistry and was a pioneer of constructive intellectual dialogue ­between science and religion.6 I began to realize that I could see my faith and my science as offering different, yet potentially complementary, ­perspectives on reality. It was a useful insight; yet I was aware that this was a staging post along the way to something better, rather than the final destination of my reflections. What Coulson offered me was indeed a helpful way of thinking. Yet I found myself dissatisfied by the prospect of merely neutralizing a potential problem—the possible conflict of science and religion, which would require me to compartmentalize these within my mind, locating each in distinct and insulated silos. I did not want a problem to be removed; I wanted a more capacious way of thinking that would hold science and faith together. The more satisfying positive solution that I developed a year or so later arose through my growing admiration for the writings of C. S. Lewis. I found myself captivated by his essay “Is Theology Poetry?,” which I first read in February 1974. Lewis here argued that Christianity set out a capacious vision of reality, capable of accommodating the natural sciences within its richly textured imaginative framework. The final sentence of that essay quickly became the focus of my emerging idea of the “mind of Christ” (1 Cor 2:16) and my awareness of the importance of a “discipleship of the mind,” in which I would aim to love God with all my mind, not merely in my heart (Mk 12:30). “I believe in Christianity as I believe that the Sun has risen, not only because I see it but because by it I see everything else.”7 Through reading and reflecting on Lewis, I began to realize that I needed to study Christian theology properly in order to grasp its intellectual riches and build bridges with the natural sciences. This, however, See especially C. A. Coulson, Science and Christian Belief (London: Oxford University Press, 1955). This work is still worth careful study. 7 C. S. Lewis, They Asked for a Paper (London: Bles, 1962), 165. 6

244

Alister E. McGrath

seemed an impossible goal. So I continued to study the natural sciences at Oxford, and went on to undertake doctoral research in the laboratories of Professor Sir George Radda in Oxford’s Department of Biochemistry. This research went remarkably well and led to me gaining a scholarship at Merton College Oxford. The unusual nature of this scholarship eventually allowed me to continue my scientific research, while at the same time taking Oxford’s undergraduate degree in Christian theology. It was a steep learning curve, but I knew it was necessary and believed it would be enriching. It was. In three years I gained my doctorate in molecular biophysics, and secured first class honors in theology. I could now begin the process of correlating my faith and science in an informed manner. I wish I had had access to a work like this collection of essays back then, which would introduce me to a range of significant and interesting Christian theologians through its perspicuous yet accessible analysis. This would have enabled me to grasp the insights of some leading thinkers so that I could then weave these into my own way of thinking about the relation of my science and faith. Readers are not being asked to agree with these ten writers; they are rather being invited to discover how they might help them in their own reflections. Some of the theologians represented here are more closely aligned with the natural sciences than others—Thomas F. Torrance and Wolfhart Pannenberg immediately come to mind. I suggest that readers see these essays as gateways into the rich and complex world of Christian theology and use them to reflect on the connections that they can make with their own interests. Yet this volume might also encourage its readers to extend their range to include others who have written on the doctrine of creation—for example, Athanasius, Augustine, Thomas Aquinas, and Jonathan Edwards, who all offer potentially interesting approaches. So why the decision to focus on the doctrine of creation in this volume? In my view, this was a wise and justified move. Although the historical question of the origins of the Scientific Revolution is widely conceded to be complex,8 there is little doubt that the Christian notion See, for example, Peter Dear, Revolutionizing the Sciences: European Knowledge and Its Ambitions, 1500–1700, 2nd ed. (Princeton, NJ: Princeton University Press, 2009); David Wootton, The Invention of Science: A New History of the Scientific Revolution (London: Allen Lane, 2015).

8

Afterword

245

of an ordered creation, endowed with a rationality grounded in its creator, played a critical role in stimulating the emergence of the natural sciences.9 Johann Kepler, one of the more theologically informed natural scientists of the seventeenth century, considered the harmony of the universe to reflect the wisdom of God its creator.10 More recently, John Polkinghorne—a British quantum theorist turned theologian—stressed the importance of humanity being created bearing the “image of God” in helping us understand the astonishing human capacity to make sense of our universe. How are we to account for the ability of mathematics to model so accurately the fundamental structures of the universe? As Polkinghorne noted, the Christian doctrine of creation affirms a fundamental “congruence between our minds and the universe, between the rationality experienced within and the rationality observed without.”11 With the exception of Craig Bartholomew’s discussion of Kuyper, the concept of the image of God does not feature prominently in this collection of essays. This is clearly an aspect of the doctrine of creation that is ripe for further exploration. The Christian doctrine of creation thus sets in place two core themes that are fundamental to the scientific enterprise—the notion of an ordered and intelligible universe, and that of a human mind with a created capacity to make sense of this universe. Yet this doctrine of creation enables a further intellectual move, which immediately distinguishes a Christian from a secular approach to the natural sciences—namely, that the close study of the natural world leads to an enhanced appreciation of the wisdom and beauty of God as its creator.12 As John Calvin and others have emphasized, there is a fundamental religious motivation for Other factors, of course, must be noted, such as Peter Harrison’s reflections on the role of original sin in stimulating the quest for a more reliable scientific knowledge: Peter Harrison, The Fall of Man and the Foundations of Science (Cambridge: Cambridge University Press, 2007). 10 Giora Hon, “Kepler’s Revolutionary Astronomy: Theological Unity as a Comprehensive View of the World,” in Tamás Demeter, Kathryn Murphy, and Claus Zittel, eds., Conflicting Values of Inquiry: Ideologies of Epistemology in Early Modern Europe (Leiden: Brill, 2014), 155-75. 11 John Polkinghorne, Science and Creation: The Search for Understanding (London: SPCK, 1988), 20-21. 12 Susan E. Schreiner, The Theater of His Glory: Nature and the Natural Order in the Thought of John Calvin (Durham, NC: Labyrinth Press, 1991). See also Peter Harrison, “Sentiments of Devotion and Experimental Philosophy in Seventeenth-Century England,” Journal of Medieval and Early Modern Studies 44, no. 1 (2014): 113-33. 9

246

Alister E. McGrath

the study of the natural sciences—a deepened sense of wonder and joy at being able to study God’s works, and appreciate the correlation with God’s words. This point lies behind the great Renaissance metaphor of the “Two Books”—the book of God’s words (Scripture), and the book of God’s works (the created order). These two books could be read side by side and allowed to enrich each other. They share the same author, yet display their insights in different manners.13 As John Calvin put it: “The knowledge of God, which is clearly shown in the ordering of the world and in all creatures, is still more clearly and familiarly explained in the Word.”14 This point highlights one aspect of the doctrine of creation that is noted at various points in this volume, particularly in Craig Bartholomew’s discussion of Kuyper, Katherine Sonderegger’s reflections on Barth, and Kevin Vanhoozer’s account of Torrance—the notion of a “natural theology.” This term is ambivalent, and is interpreted in a manner of ways, including proving the existence of God through an appeal to either the natural order, or using “natural” human faculties, unaided by divine revelation.15 Yet despite these difficulties of interpreting this notion, the idea that what God creates can tell us something about God as creator remains theologically and apologetically significant. Many natural scientists intuitively develop their own idiosyncratic concepts of natural theology in trying to articulate their sense that a complex and beautiful universe somehow helps them grasp more fully the wisdom and beauty of God. Perhaps the reflections on this theme contained in this volume may help them clarify and extend these ideas. The ten theologians here considered each offer a distinct account of the importance of the doctrine of creation. Some emphasize the importance of creation within an overall Christian worldview (Kuyper); others place their emphasis elsewhere. Some are clearly shaped by cultural Kenneth J. Howell, God’s Two Books: Copernican Cosmology and Biblical Interpretation in Early Modern Science. Notre Dame, IN: University of Notre Dame Press, 2002. 14 John Calvin, Institutes of the Christian Religion, I.x.1. See further Stephen J. Grabill, Rediscovering the Natural Law in Reformed Theological Ethics (Grand Rapids, MI: Eerdmans, 2006), 70-97. 15 For reflections on this notion, see Russell Re Manning, ed., The Oxford Handbook of Natural Theology (Oxford: Oxford University Press, 2013); Alister E. McGrath, Re-Imagining Nature: The Promise of a Christian Natural Theology (Oxford: Wiley-Blackwell, 2016). 13

Afterword

247

debates (e.g., Warfield, in response to Darwinian theories of evolution). Yet perhaps most importantly, these ten figures offer us helpful insights into how we might explore connections between the natural sciences and faith. In her insightful essay on Barth, for example, Katherine Sonderegger challenges the widely held view that Barth saw science and theology as having little to say to each other. Barth’s vision of the doctrine of creation embraces the totality of reality, including the realm of the natural sciences. Although Barth might not resolve the question of how science and religion might have a meaningful conversation, he can be seen to set a helpful context for such a discussion. As the editors point out, the doctrine of creation can be seen as a “distributed” (John Webster) notion, interlinking and interconnecting with other areas of theology.16 We might think, for example, of salvation in terms of the restoration or renewal of God’s work of creation.17 Or, to note an area of discussion that is touched on at several points in this volume, in what ways can we see creation as disclosing the nature of Christ?18 To study the doctrine of creation is thus to begin a journey of exploration that leads to a deepened understanding of other areas of Christian thought. Many theologians now tend to think of Christian theology as a whole as a web of interconnected ideas, rather than as individual isolated compartments of thought. The doctrine of creation is thus an excellent starting point for exploring the fundamental intellectual coherence of Christianity, opening up the fascinating question of how this remarkable “Big Picture” of reality is able to accommodate the natural sciences, explaining their successes and at the same time identifying their limits. This volume will help its readers Webster sees creation as “one of two distributed doctrines in the corpus of Christian dogmatics,” the other such distributed doctrine being the Trinity, “of which all other articles of Christian teaching are an amplification or application, and which therefore permeates theological affirmations about every matter.” See further John Webster, God Without Measure: Working Papers in Christian Theology, Vol. 1: God and the Works of God (London: Bloomsbury T&T Clark, 2016), 115–26; quote at p. 117. 17 Brenda B. Colijn, Images of Salvation in the New Testament (Downers Grove, IL: IVP Academic, 2010), 102-20. 18 This theme is significant in the writings of Jonathan Edwards: see, for example, Belden C. Lane, “Jonathan Edwards on Beauty, Desire and the Sensory World,” Theological Studies 65, no. 1 (2004): 44-68. Edwards here draws on the prologue to John’s gospel, which speaks of Christ as the agent of creation (Jn 1:3). 16

248

Alister E. McGrath

gain a sense of the deeper vision of reality that the Christian faith enfolds and enables, which is able to affirm the importance of the sciences, while guarding against their misinterpretation.19 There is much more that needs to be said about the relation of the natural sciences and the Christian faith than the doctrine of creation—but it is a wonderful place from which to start.

Many have expressed concern about the intellectual arrogance of “scientism” (the view that the sciences alone offer reliable and relevant knowledge), or the impoverishing results of excessive reductionism. For some representative engagements with such concerns, see Jeroen de Ridder, Rik Peels, and René van Woudenberg, eds., Scientism: Prospects and Problems (New York: Oxford University Press, 2018); Steven W. Horst, Beyond Reduction: Philosophy of Mind and PostReductionist Philosophy of Science (Oxford: Oxford University Press, 2007).

19

C O N T R I BU T O R S

Craig Bartholomew (PhD, Bristol University) is the director of the Kirby Laing Institute for Christian Ethics at Tyndale House, Cambridge. Formerly he was senior research fellow at the University of Gloucestershire and recently the H. Evan Runner Professor of Philosophy and Professor of Religion and Theology at Redeemer University College in Ancaster, Canada. He has edited and written many books, most recently The Doctrine of Creation: A Constructive Kuyperian Approach (IVP Academic, 2020), coauthored with Bruce Riley Ashford; The God Who Acts in History: The Significance of Sinai (Eerdmans, 2020); and Contours of the Kuyperian Tradition: A Systematic Introduction (IVP Academic, 2017). Bradley Gundlach (PhD, University of Rochester) is Distinguished Professor of History at Trinity International University. He serves as director of the division of humanities and as book review editor for Fides et Historia, the journal of the Conference on Faith and History. Dr. Gundlach is the author of Process and Providence: The Evolution Question at Princeton, 1845–1929 (Eerdmans, 2013), and is currently at work on a biography of Princeton theologian B. B. Warfield. Joshua W. Jipp (PhD, Emory University) is associate professor of New Testament at Trinity Evangelical Divinity School. He is the author of Divine Visitations and Hospitality to Strangers in Luke-Acts: An Interpretation of the Malta Episode in Acts 28:1-10 (Brill, 2013), Christ Is King: Paul’s Royal Ideology (Fortress Press, 2015), Saved by Faith and Hospitality (Eerdmans, 2017), and Reading Acts (Cascade, 2018). His most recent monograph is The Messianic Theology of the New Testament (Eerdmans, 2020). He also currently serves with Robert Yarborough as coeditor of the Baker Exegetical Commentary on the New Testament.

250 Contributors

Alister E. McGrath (DPhil, Oxford University) is Andreos Idreos Professor of Science and Religion at Oxford University and is the director of the Ian Ramsey Centre for Science and Religion. McGrath is a prolific author and speaker. He has written almost one hundred articles and nearly forty books, including A Scientific Theology (3 vols.; T&T Clark, 2001–2003) and The Dawkins Delusion? Atheist Fundamentalism and the Denial of the Divine (SPCK, 2007). McGrath has also been invited to present for prestigious academic lectures, including the Bampton, Gifford, and Boyle Lectures. He is also a founding member of the International Society for the Study of Science and Religion and has received several honorary doctorates. Murray Rae (PhD, Kings College, London) is professor of theology at the University of Otago, New Zealand. Rae was first trained as an architect, and his most recent publication, Architecture and Theology: The Art of Place (Baylor University Press, 2017) investigates how the art forms engaged in the construction of our built environment relate to Christian faith. His other publications include Christian Theology: The Basics (Routledge, 2015), Kierkegaard and Theology (T&T Clark, 2010), History and Hermeneutics (T&T Clark, 2005), and Kierkegaard’s Vision of the Incarnation: By Faith Transformed (Oxford Clarendon Press, 1997). Fred Sanders (PhD, Graduate Theological Union) is professor of theology in the Torrey Honors Institute at Biola University, where he has taught since 1999. He is the author or editor of over a dozen books, most recently Third Person of the Trinity: Explorations in Constructive Dogmatics (Zondervan, 2020); with Oliver Crisp, Retrieving Eternal Generation (Zondervan, 2017); with Scott Swain, The Triune God in Zondervan’s New Studies in Dogmatics Series (Zondervan, 2017); and John Wesley on the Christian Life: The Heart Renewed in Love (Crossway, 2013). Along with Oliver Crisp, Sanders cofounded the annual LA Theology Conference, and he blogs regularly at Scriptorium Daily. Christoph Schwöbel (PhD, Philipps Universität Marburg) is professor of systematic theology at St. Mary’s College, University of St. Andrews, Scotland. His most recent publications include Die Stadt: Interkulturelle theologische Zugänge; edited with Jürgen Kampmann, Luther Heute:

Contributors

251

­ usstrahlungen Der Wittenberger Reformation (Mohr Siebeck, 2017); Der A entgrenzte Kosmos und der begrenzte Mensch: Beiträge zum Verhältnis von Kosmologie und Anthropologie (Vandenhoeck & Ruprecht, 2016); and Word and Spirit: Renewing Christology and Pneumatology in a Globalizing World (De Gruyter, 2014). Katherine Sonderegger (PhD, Brown University) holds the William Meade Chair in Systematic Theology at Virginia Theological Seminary. Her most recent publication is Systematic Theology, Volume 2: The Doctrine of the Holy Trinity: Processions and Persons (Fortress Press, 2020), which follows her widely acclaimed Systematic Theology: Volume 1, The Doctrine of God (Fortress Press, 2015), the first of a three-volume dogmatics, described by John Webster as “one of the most distinguished treatments of the Christian doctrine of God in recent decades.” Her essays have appeared in the Scottish Journal of Theology, Harvard Theological Review, Zeitschrift für Neuere Theologiegeschichte, and the International Journal of Systematic Theology. Kevin Vanhoozer (PhD, Cambridge University) is research professor of systematic theology at Trinity Evangelical Divinity School. His most recent publications include Hearers and Doers: A Pastor’s Guide to Making Disciples Through Scripture and Doctrine (Lexham Press, 2019) and Biblical Authority after Babel (Brazos Press, 2016). His two volumes The Drama of Doctrine (Westminster John Knox Press, 2006) and Faith Speaking Understanding (Westminster John Knox Press, 2014) were both awarded the Christianity Today Best Theology Book of the Year Award. He also serves as editor with Daniel Treier of the Studies in Christian Doctrine and Scripture series with IVP Academic. Stephen N. Williams (PhD, Yale University) is professor of systematic theology at Union Theological College in Belfast, Northern Ireland and was senior research fellow with the Creation Project for the 2017–2018 academic year. He studied modern history at Oxford and theology at Cambridge and gained his doctorate through the Department of Religious Studies at Yale University. He has published on modern theological epistemology in Revelation and Reconciliation (Cambridge University Press, 1996) and on Nietzsche’s critique of Christianity in The Shadow of

252 Contributors

the Antichrist (Baker Academic, 2006). He has also delivered the 2009 Kantzer Lectures, published as The Election of Grace: A Riddle Without a Resolution? (Eerdmans, 2015). Stephen John Wright (PhD, Charles Strut University) is lecturer in Christian theology at Nazarene Theological College, Manchester. He is the author of Dogmatic Aesthetics: A Theology of Beauty in Dialogue with Robert W. Jenson (Fortress Press, 2014) and the editor of Robert W. Jenson, Theology as Revisionary Metaphysics: Essays on God and Creation (Cascade, 2014). Together with Chris E. W. Green, he has edited The Promise of Robert W. Jenson’s Theology: Constructive Engagements (Fortress Press, 2017).

GENERAL INDEX activity, 65-73, 81-82, 94-95, 135, 215 See also providence Adam, 10, 79, 193, 206, 209 creation of, 62, 227 historical, 78, 80 agency communicative, 221 divine and creaturely, 33, 236-38 and God, 33, 135, 176-78, 185, 198, 200, 220, 229 human, 72, 199 personal and impersonal, 218 supernatural, 65 See also intervention; providence Alexandria, 122, 129 analogy of being, 102, 138 Aquinas, 140, 150, 244 and Scholasticism, 137-38, 178 Aristotelianism, 2, 10, 110, 129, 131, 139, 140, 178, 197, 202 See also causality; logic Athanasius, 7, 9, 124, 128-29, 141, 244 atheism, 17-8, 64, 170, 242 materialist, 23 modern, 218 new, 242 Barth, 121-23, 129-30, 138, 145, 170, 173, 193, 215, 246-47 Bible authority of, 66, 72 and creation, 23, 27, 79, 81, 89, 97-98 and interpretation, 16, 26, 28, 67-68, 94, 96, 102, 143 and science, 63, 78, 94 and theology, 46 use of, 26-27, 30 biblicism, 48, 57

causality, 126-27, 178, 182, 186-87, 217, 219, 229 fourfold, 186, 197 and God-creation relationship, 33, 73, 108, 110, 117, primary and secondary, 200 See also agency; Aristotelianism Christ, 69, 72, 77, 79-80, 123, 135, 145-47, 165, 177 and creation, 9, 104-8, 121-22, 125-31, 138-39, 161, 203, 211, 228-30, 235-36, 247 communication and God, 128, 135, 159, 185-89, 209 consciousness, 40-42, 52, 119 religious 107-10 See also piety contingency, 9, 11, 171-90, 208, 214, 223-24, 226 of creation, 90, 134-39, 144-46, 159 See also freedom of God; necessity cosmology, 2, 25-26, 89, 96-97, 106, 111-13, 210, 223 creation days of, 72-74, 106, 116 doctrine of, 1-9, 18-19, 30-32, 38, 73-74, 97-98, 99-100, 104-18, 150-52, 157, 167, 173-74, 193, 197, 203, 211, 213-16, 220-28, 232-34, 237-40, 244-48 and eschatology, 3, 106, 153-62, 174, 177, 183, 190, 227-28 ex nihilo, 2, 9, 22, 62, 71-73, 81, 91, 110, 134-37, 154-61, 222-29 goodness of, 2, 42, 238

mediate, 9, 70-73, 80-82 primary and secondary, 4, 18-33 and teleology, 97, 227, 239 Darwinism, 16, 51, 54, 61, 63-66, 204, 219 Social, 49 Dawkins, Richard, 218-19, 242 deism, 66, 70, 73, 175, 198, 218, 228-29 demythologization. See hermeneutics: and demythologization Descartes, 8, 10, 125, 179, 185, 216-17 design, 50, 53, 56, 64-65, 204 developmentalism, 5, 62, 76-77 dialectic theology, 86, 102 divine attributes, 18-20, 32 omnipotence, 19-21, 72, 183, 230 omnipresence, 160 omniscience, 161 dualism, 23, 74, 86-87, 92, 125-28, 132, 135-40, 217 science and religion, 48, 51, 57 dynamis and theiotes, 21 Edwards, Jonathan, 193, 197-99, 206 Einstein, Albert, 7, 9, 122-33, 138-45 empiricism and science, 44, 59, 141, 224, 226, 232, 234, 239 Enlightenment, 10, 36-40, 50, 55, 197, 214-18, 220, 241 epistemology, 7, 54-6, 122, 125-28, 141, 194-95, 215-17, 221 eschatology, 3, 94, 106, 148, 153-62, 183, 190, 227-28 ethics, 49, 107, 159, 216, 239 and evolution, 52 event, 114-16, 176-77, 183, 186, 190

254 evidentialism, 28-29 evolution, 2, 5, 8, 9, 48-49, 63, 113, 152, 187 and Christianity, 11, 16, 24, 38, 42, 50-56, 64-77, 162-64, 214, 231 and human origin, 4. 10, 26-8, 59, 77-82, 204-6, 238 materialistic, 24, 26, 82 theistic, 59-63, 82 theory and worldview distinction, 35, 219 faith and science, 19, 41-42, 47-48, 51, 57, 64, 86, 92-97, 242-44, 247 fall, 78-79, 228 and world, 42, 54, 136, 237 See also sin Faraday, Michael, 8, 10, 181-85, 226, 233-34 fideism, 28-30 form and matter, 20, 24-26, 30-31 freedom of God, 19-21, 66, 106, 113, 134-36, 144, 159, 178, 200, 207, 223, 230 and human beings, 110-11, 159, 215, 219, 227-28, 231 See also contingency; necessity fundamentalism, 59, 61, 64, 69-70, 82 Genesis, 2, 4, 21, 25-30, 48, 72, 89, 100, 104, 109-17, 162, 188, 192, 195-96, 204, 209, 223-24, 231-37 God. See Christ: and creation; communication and God; divine attributes; freedom of God; Holy Spirit; imago Dei; providence; Trinity Gould, Stephen Jay, 112, 133 grace, 14, 75, 77, 104, 221 common, 5, 36, 38-41 Haeckel, Ernst, 10, 49-51 Harrison, Peter, 11, 225, 233 Hawking, Stephen, 8-9, 201-3 hermeneutics, 56, 162-67, 186 and demythologization, 86-9, 92-94, 97

General Index historical criticism, 16 history, 43, 51, 60, 63-64, 155, 199 and God, 33, 71, 74-76, 91, 127, 145, 157-58, 185, 193, 195, 197, 199-201, 207 salvation, 113, and theology, 2, 7-8, 10-11, 82, 84-85, 104-19, 153-54, 162, 165, 167, 171, 176-77, 188, 190 See also saga Holy Spirit, 14, 28-29, 115, 123, 130, 134, 139, 151, 156-62, 165, 188-90, 200, 214, 220, 228-31, 234-39 and field of force, 172, 181, 185 homoousion, 7, 128-41 Hume, David, 126, 236-37 idealism, 16, 101 imaginary and real time, 8, 201-3 imago Dei, 5, 26, 38-40, 46, 96, 139, 204, 245 inerrancy, 59-61, 72 intelligibility, 2, 127-37, 145-46, 202, 222-26, 232, 245 See also order of creation intervention, 66, 108, 179-80, 199-200, 207 See also agency; providence Irenaeus, 227, 231 Kant, Immanuel, 10, 37, 101, 109, 124, 126-27, 136, 186, 193, 202, 216 kataphysics, 7, 123-24, 128-33, 136-46 See also poetics kenosis and creation, 161 kerygma, 94, 97 knowledge and God, 76, 89-90, 94, 123, 128, 137-39, 214, 221, 231, 245-46 archetypal and ectypal, 45-48 Leibniz, Gottfried Wilhelm, 10, 168, 175-76, 180 liberalism, 64, 68-69, 82, 86, 101-2

See also fundamentalism logic, 43, 47, 95, 132, 138-39, 235-37 Aristotelian, 129-31 Logos, 39, 46, 53, 134-36, 147, 189, 235-37 materialism, 17-8, 23-26, 82, 106, 194, 199, 204 Maxwell, James Clerk, 7, 10, 130-31, 144, 185 metaphor, 57-8, 184, 197-99, 207, 234, 246 metaphysics, 3, 4, 10, 19, 24, 30, 63-64, 94-96, 110-17, 123, 125, 136-40, 159-60, 181, 197-99, 202, 205, 206-7, 218 miracle, 42-43, 50, 65-71, 78, 85, 88, 95-96, 133, 236-37 See also wonder myth, 6, 86-97, 113, 117-18 narrative, 8, 114, 117-18, 145, 196, 201, 209 and creation, 2, 4, 105-10, 116, scientific, 8, 195, 242 natural law, 62, 67, 73, 95, 112, 173-76, 228, 238 See also necessity natural selection, 53, 64-65, 234 natural theology, 2, 8, 45, 102, 123, 130, 137-40, 145, 152, 171, 173-77, 181, 184, 188, 246 naturalism, 44, 47, 68-9, 107, 111 and evolution, 42, 54, 75 and methodology, 65 necessity, 19, 25, 159, 172, 175-76, 178, 230 See also contingency; freedom of God Newton, Isaac, 8-10, 119, 126-27, 130-31, 135, 139-45, 150, 155, 168, 180-81, 197, 201, 219 nihilism, 52, 211 objectification, 86-7, 93-94 ontology, 2, 6, 54, 86, 94-7, 122, 125, 128, 132, 141, 193, 196-97, 215, 217, 228 open system, 7, 153-54, 157-62,

255

General Index 166 order of creation, 2, 40, 44, 52, 56, 117, 130, 134-44, 147, 153, 158-59, 162, 179, 198, 223, 225-26, 230-32, 235-39, 246 origin, 21-22, 39, 65, 75, 89-90, 96, 101, 106-8, 116-19, 174, 183-84, 219 human, 24, 26-27, 61, 80, 205 soul, 66 See also creation palingenesis, 38-47, 57 panentheism, 200 pantheism, 18, 23-24. 160 personalism, 10, 114, 218 phenomena, 22-24, 42-44, 142, 189, 198, 226 and noumena. 127, 136 Philoponus, John, 7, 9, 122, 129, 223-24 philosophy of science, 50, 56-57, 171 piety, 106-12, 127 Platonism, 31, 37, 107, 109, 125, 196, 224-25 poetics, 117, 124-40, 145-46 Polanyi, Michael, 10, 141-42, 217, 221-22 polydimensional reality, 188-90 polytheism, 23-24 positivism, 17 prayer, 193, 200, 205-7 presupposition, 28, 56, 172-73, 177-80, 185, 189, 199 Protestantism, 101, 150 providence, 6, 18-20, 33, 62, 69-75, 80-82, 109, 156, 193-94, 201, 207 See also agency; intervention

purpose, 2, 39, 69, 105, 118-19, 136, 216, 223, 227-29, 235-36, 239 rationalism, 29, 75, 214, 241 and science, 217, 245 realism, 39-40, 56, 102, 112, 122, 132, 136-37, 140-41, 144-45 reason, 27-9, 46, 72, 123, 127, 136, 163-64, 214, 216, 220 religious consciousness or feeling, 106-11 See also piety resurrection, 69, 105-6, 188, 199, 205, 214-15, 229, 236 revelation, 5, 8, 18-23, 27-32, 38-9, 44-7, 56, 76, 88, 92, 94-98, 127, 130, 133, 138-39, 142-43, 163-64, 171, 185, 188, 220-21, 246 Sabbath, 151, 157-58, 162, 175 saga, 6, 100, 105-6, 113, 116-19 Scholasticism, 137-38 See also natural theology scientific method 1, 95, 122-23, 128-30, 196, 210, 240 scientism, 118, 219, 234, 240 Scotus, Duns, 9, 178 sin, 36, 38-43, 52-6, 62, 79, 86, 90, 97, 116, 228, 238 See also fall sovereignty, 21, 33, 57, 106, 113, 222-23 See also agency, providence space and time, 109, 121, 127-32, 135-36, 147, 152, 155, 161, 181-85, 203, 227 Spencer, Herbert, 10, 48-49 Spinoza, 23, 110, 179, substance, 2, 24, 26, 69, 122, 125, 128, 197-98 supernaturalism, 6, 65, 68-70, 75, 77

teleology, 97, 223, 227, 239, 246 See also purpose temporality, 105-6, 119, 128, 145, 154-55, 201 of creation, 116-17 and God, 33, 110, 190, 203 and history, 85, 118 and timelessness theism, 78, 96, 138 theology dialectic, 84, 86, 99, 102, natural, 2, 45, 102, 123, 130, 137-40, 145, 246 of nature, 2, 8, 137-39, 152, 171, 173, 177, 181, 184, 188 and science, 1-10, 35, 47, 68, 86, 97, 100-101, 111, 113, 118, 121, 124, 145, 149, 152-53, 162-68, 186, 192, 194-95, 206, 208, 210-11, 213, 221, 229, 232-35, 240, 247 transcendence, 31-33, 135-36, 156-61, 180, 193, 196 transmutation, 63-66, 69, 76, 78 Trinity, 7, 9, 128-31, 138, 149, 156, 159-61, 182, 188, 190, 213, 219, 233, 240 wisdom, 7, 19-21, 25, 32, 39-40, 46, 71, 144, 147, 153-55, 163, 167, 230-31, 239, 245-46 Wissenschaft, 35, 38 Wollebius, Johannes, 9, 67, 70-73, 82 wonder, 88, 95, 211 of science, 132 See also miracle worldview, 5, 35-41, 51, 53-56, 69, 80, 86, 93, 97, 118, 246 zimzum, 160, 163, 167-68

SCRIPTURE INDEX Old Testament Genesis 1, 21, 25, 100, 104, 109, 192, 195, 196, 223, 224, 235, 237 1–2, 25, 66, 109, 110, 116 1–3, 89, 109 1–5, 188 1:1, 31 1:1–2:4, 48 1:5, 31 1:16, 31 1:21, 31 1:25, 31 1:26, 139 1:27, 31 1:28, 39 2, 81 2:3, 31 2:7, 231 2:19-20, 232 3, 116

Job 33:4, 231 Psalms 19:1, 139 104:29, 91, 231 104:30, 90, 231 Proverbs 1:7, 40 8:22-31, 39 Isaiah 24:7, 236 45:9-12, 90 64:8, 90 Ezekiel 37:10, 231 New Testament

Exodus 33:23, 133

Matthew 1:18, 238 19:26, 21 19:28, 39 24:40, 144

Deuteronomy 4:19, 211

Mark 12:30, 243

Luke 1:35, 238 1:38, 238 John 1, 21, 39 1:1-2, 134 1:1-3, 92, 235 1:3, 134, 139, 247 1:3-4, 235 1:12-13, 92 1:14, 133, 189, 231 2:1-11, 236 3:1-16, 92 3:36, 92 4:24, 181, 182 5, 236 6, 236 10:10, 235 Romans 1, 21 8, 157 8:2, 231 8:11, 231 8:23, 183 1 Corinthians 2:16, 243 8:5-6, 91

8:6, 91 12:7-11, 231 2 Corinthians 3:6, 231 Galatians 3:16, 145 5:22, 231 Ephesians 1:10, 145 1:11, 52 Philippians 2, 161 Colossians 1, 21 1:15-16, 122 1:15-20, 235 1:16-17, 135 Titus 3:5, 39 Hebrews 11, 22 11:27, 22

P R A I SE F O R S C I E N C E A N D T H E D O C T R I N E OF C R E AT I ON

“This volume represents another valuable contribution of the Creation Project to our understanding of this vital doctrine. The thinkers profiled are influential, and the chapter authors are insightful. As the editors suggest, these case studies frequently deepen our awareness that seeking appropriate concord between theology and science is complex but inevitable for biblical Christians.” Daniel J. Treier, Knoedler Professor of Theology at Wheaton College “The person who does not specialize in the work of modern academic theologians wants to know who the key writers are, with a clear survey of their distinctive views and contributions, given with sympathy and even critique. This volume, with its focus on the specific question of how these theologians have brought the Christian doctrine of creation into engagement with the sciences, has achieved exactly that. The editors, representing the Carl Henry Center for Theological Understanding, have done great service to us all and furthered the invaluable work of the center. We owe them a deep debt of gratitude!” C. John (“Jack”) Collins, professor of Old Testament at Covenant Theological Seminary, St. Louis “How did the most important twentieth-century Protestant theologians model Christian engagement with the scientific questions of their time? Rather than staking out a definitive position on creation and science, each of the ten essays in this book gives an account of the diverse ways a particular theologian (Warfield, Barth, Torrance, Moltmann, Pannenberg, etc.) addressed current scientific issues and how he understood the very relationship of science and theology in light of the doctrine of creation. This is a rich feast indeed!” J. Richard Middleton, professor of biblical worldview and exegesis, Northeastern Seminary at Roberts Wesleyan College, Rochester, New York

“Science and the Doctrine of Creation presents ten of the most influential nineteenth- and twentieth-century theologians writing on dialogue with the sciences, analyzed by ten leading contemporary scholars in the field. With this new book, Geoffrey Fulkerson and Joel Chopp confirm the leading position of the Carl F. H. Henry Center for Theological Understanding as a forum for informed scholarly debate, bringing theology and science into fruitful interaction.” Lydia Jaeger, lecturer and academic dean at the Institut Biblique de Nogent-sur-Marne, France, and research associate at St. Edmund’s College, University of Cambridge “For too long Christians have weaponized the seeming conflict (or harmony) between science and theology. Good historians urge us instead to pay attention to the actual practice of particular theologians and scientists: ‘Don’t generalize in the abstract; look and see!’ they say. Science and the Doctrine of Creation is a wonderful example of this salutary approach, drawing on theological luminaries like Barth, Torrance, and Pannenberg as guides into the rich meaning of creation. These learned essays remind us that the dogmatic issues surrounding God’s creation extend far beyond the customary debates over origins. Readers will naturally gravitate to some chapters over others, but the ten chapters taken together offer a banquet of stimulating analysis for famished readers. What a welcome addition to the science and theology dialogue!” Hans Madueme, associate professor of theological studies at Covenant College in Lookout Mountain, Georgia

A B O U T T H E AU T HO R

Geoffrey H. Fulkerson (PhD, TEDS) is assistant director of the Carl F. H. Henry Center for Theological Understanding at Trinity Evangelical Divinity School. He is founder and editor-in-chief of HCTU’s periodical, Sapientia.

Joel Thomas Chopp (PhD candidate, University of Toronto) is project and communications manager for the Henry Center’s Creation Project. He also serves as associate editor of Sapientia.

Please visit us at ivpress.com for more information about Geoffrey Fulkerson and Joel Chopp and a list of other titles they’ve published with InterVarsity Press.

Discover more titles from

InterVarsity Press

Click to view the newest and trending titles in Academic Texts & Reference IVP Academic covers disciplines such as theology, philosophy, history, science, psychology, and biblical studies with books ranging from introductory texts to advanced scholarship and authoritative reference works. Culture, Mission, and Christian Life Our books are deeply biblical and profoundly practical, discussing topics like Christian spirituality, prayer, evangelism, apologetics, justice, mission, and cultural engagement. Bible Studies & Group Resources IVP provides Bible studies and small group resources for you and your church, helping individuals and groups discover God’s Word and grow in discipleship. Spiritual Formation Formatio books follow the rich tradition of the church in the journey of spiritual formation. These books are not merely about being informed, but about being transformed by Christ and conformed to his image. Church Leadership IVP Praxis brings together theory and practice for the advancement of your ministry using sound biblical and theological principles to address the daily challenges of contemporary ministry.

Click below to view more books in these categories Apologetics

Discipleship

Philosophy

Biblical Studies

Family, Children & Youth

Psychology

Career & Vocation

Fiction

Race & Ethnicity

Church & Culture

Justice/Peace

Science

Church History

Spiritual Formation

Theology

Commentaries

Missions & Missiology

Youth Ministry

For a list of IVP email newsletters please visit  ivpress.com/newsletters.

Sign up for Books & Deals