<<

PROVIDENCE AND -: RETHINKING GOD’S RELATION TO THE

WORLD THROUGH THE EYES OF

Dissertation

Submitted to

The College of Arts and of the

UNIVERSITY OF DAYTON

In Partial Fulfillment of the Requirements for

The Degree of

Doctor of in

By

John Forrest Birch, M.S., M.A.

UNIVERSITY OF DAYTON

Dayton, Ohio

December, 2020

PROVIDENCE AND SPACE-TIME: RETHINKING GOD’S RELATION TO THE

WORLD THROUGH THE EYES OF JOHN POLKINGHORNE

Name: Birch, John Forrest

APPROVED BY:

Daniel S. Thompson, Ph.D. Faculty Advisor

Brad J. Kallenberg, Ph.D. Faculty Reader

Vincent J. Miller, Ph.D. Faculty Reader

Dennis M. Doyle, Ph.D. Faculty Reader

Allen J. McGrew, Ph.D. Outside Faculty Reader

______Jana M. Bennett, Ph.D. Chairperson

ii

© Copyright by

John Forrest Birch

All rights reserved

2020

ABSTRACT

PROVIDENCE AND SPACE-TIME: RETHINKING GOD’S RELATION TO THE

WORLD THROUGH THE EYES OF JOHN POLKINGHORNE

Name: Birch, John Forrest University of Dayton

Advisor: Dr. Daniel S. Thompson

Recent have witnessed a greater acknowledgement among physicists and theologians that theology and have something to say to each other with regard to the endeavor to better understand God’s relationship to the world. Modern physics has been led by contemporary scientific pioneers who saw unique ways to solve problems that began to arise as more and more aspects of physical reality were investigated and required explanation. The insights arising from these physicists served to banish the older traditional scientific view of the as a “grand machine”. One of the results of this modified scientific view of the nature of the universe has been renewed dialog between physics and theology in a new with a whole new set of challenges to longstanding questions about God’s relation to the world.

One of the major participants in this discussion is John Polkinghorne, a high- energy particle physicist and Anglican , who approaches the pursuit of knowledge by beginning with physics and proceeding to examine the realm of natural phenomena for evidence of agreement and consistency with the claims of religious faith. Polkinghorne’s view seems to offer promise because of 1) his effort to remain true to the tenets of the

Christian faith, and 2) his reliance on sound modern and mathematics (in

iii particular quantum indeterminacy and ). The promise Polkinghorne offers is the use of intellectual pursuit, by way of modern physics, in order to facilitate faith’s search for understanding with regard to the question of God’s relationship with the world.

This can have implications not only for academic theology, but also for ’ grasp of various religious concepts such as creation, , , and the nature of God, all of which are key doctrines believers deal with each .

Polkinghorne has said of his own endeavors in this area “My concern is to explore to what extent we can use the search for motivated understanding, so congenial to the scientific mind, as a route to being able to make the substance of Christian orthodoxy our own.”1 During my research I found that one of the key criticisms of Polkinghorne is that determinate rules underlying chaos theory undermine his reliance on indeterminacy for his view of divine interaction with the world. I am not convinced that the way he uses chaos theory undermines his reliance on indeterminacy, since indeterminacy is built into the warp and woof of reality according to the Copenhagen Interpretation of . However, my concern here is not to argue for the particular way he uses chaos theory but rather to argue that the criticisms of his use of chaos theory can better contribute to the discussion of God’s relation to the world were they to refocus their attention on the relationship between the quantum and macroscopic levels of physical reality. With the increased research efforts of modern physicists to better understand the quantum level of reality it seems that this is the area that goes most directly to the heart of the question of God’s interaction with and influence in the world.

1 John Polkinghorne, Faith of a Physicist ( Minnesota: Fortress Press, 1994), 1 iv

Finally, in making this argument I hope that it will be apparent why it is so important for the physicist, theologian, or any believer to have a better understanding of

God’s relationship with the world. I believe the fact that this question has long been pondered is evidence enough that it is fundamentally important. Contemporary challenges to models of understanding the relationship of God and world have sometimes resulted in flawed ways of trying to arrive at answers, such as the "God of the Gaps" approach of using epistemological gaps as a warrant to argue for the presence and work of God. In particular, by exploring Polkinghorne's view and how he is criticized, I hope to show that his approach helps to avoid such pitfalls, which is crucial for articulating a better understanding of the question of divine interaction with the world that is faithful to the claims of modern physics as well as Christian .

v

Dedicated to my parents and those drawn to the grandeur of the universe, and the glory of the One who made it

vi

ACKNOWLEDGEMENTS

I would like to extend my deep appreciation to Dr. Daniel S. Thompson, my dissertation advisor, for his patient guidance and advice throughout the writing of this dissertation. His guidance in helping me to keep from straying too far from my topic has been invaluable. I am also indebted to Dr. Terry Tilley for his guidance and direction as my original advisor prior to his departure for Fordham University. He helped me immensely as I worked through my general exams on my way to the qualifying exam.

Also, I wish to thank Dr. Bad J. Kallenberg for his advice and guidance as one of my professors when I first entered the program leading to the Ph.D. in theology here at the

University of Dayton.

Special thanks must also be extended to my wife Debbie, and my immediate family members for their patience and support during the that I remained “sealed away” in my office reading and writing. I am forever grateful to my parents who always supported me in my academic pursuits, and gave me the encouragement unique to loving parents who believed that I could do anything that I set my mind to. I miss them every day of my .

vii

TABLE OF CONTENTS

ABSTRACT iii

DEDICATION vi

ACKNOWLEDGEMENTS vii

CHAPTER 1 THE BIRTH OF MODERN SCIENCE AND THE QUESTION OF GOD’S RELATION TO THE WORLD 1

1.1 Viewing Physical Reality as a Great Machine 10

1.2 Losing the Mechanical View of Physical Reality 14

CHAPTER 2 MODERN PHYSICS’ VIEW OF REALITY – NEW CONTEXTS FOR OLD QUESTIONS 23

2.1 The Loss of the Constancy of Space and Time in Physics 28

2.2 The Loss of Determinacy in the Fabric of Space-Time 33

2.3 God and 38

CHAPTER 3 APPROACHING A VIEW OF GOD’S RELATIONSHIP TO THE WORLD 44

3.1 Theological Models of God’s Relation to the World 48

3.2 John Polkinghorne and the Communicator of Information Model 56

CHAPTER 4 CRITICISMS OF POLKINGHORNE’S “COMMUNICATOR OF INFORMATION” MODEL 82

4.1 Wildman’s Criticism of Polkinghorne 83

4.2 Saunders’ Criticism of Polkinghorne 91

CHAPTER 5 EVALUATING CRITICISMS OF POLKINGHORNE 102

CHAPTER 6 CONCLUDING THOUGHTS 124

BIBLIOGRAPHY 132

viii

CHAPTER 1

THE BIRTH OF MODERN SCIENCE AND THE QUESTION OF GOD’S RELATION

TO THE WORLD2

Generally it is recognized that while roots of physics, once referred to as natural

philosophy, can be traced back into antiquity, the advent of modern physics occurred in

the seventeenth . By this time Nicholas Copernicus (1473-1543) had introduced

the world to a heliocentric (-centered) solar in his published work De

Revolutionibus. The research of (1571-1630) and (1564-

1642), based on empirical information, led to statements of physical principles governing

the of objects. This included those objects of the celestial realm, a topic of

particular interest to humanity. Galileo’s use of scientific reasoning was regarded by

Albert Einstein (1879-1955) as “…one of the most important achievements in the

of modern thought and marks the real beginning of physics.”3 Galileo’s approach

demonstrated that intuitive conclusions drawn from immediate observation are not

necessarily trustworthy. Rather than relying on questions of purpose and final cause, his

approach involved determining measurable relationships between phenomena

in nature. This approach was contrary to that of Aristotle (385-323 BC). Galileo

effectively overthrew a long held notion that with respect to the laws governing natural

phenomena one should rely on the intuitive explanation as being he correct one. As an

example, with regard to motion he concluded that it is not a body’s velocity that indicates

2 I use the “world” herein to refer to nature or the universe. This is in keeping with terminology used in various scientific and theological texts. 3Albert Einstein, Leopold Infeld, The of Physics: From Early Concepts to Relativity and Quanta (New York, : Simon and Schuster, 1966), 6. 1

whether or not external influences, or forces, act on that body (the traditional intuitive

perception), but rather it is a body’s change in velocity, or . This was

articulated later by (1642-1727) in his first two laws of motion. Also

during this time Galileo’s contemporary, Francis Bacon (1561-1626), proposed a revised

system of thought with regard to the scientific endeavor. Bacon advocated for

mathematics, not as a separate science, but rather as an instrument of physics that could

be used for modeling various phenomena in nature involving material objects in motion.

In his words “…inquiries into nature have the best result when they begin with physics

and end in mathematics.”4 Therefore, probable conclusions could be reached based on

available evidence that could be measured and quantified. This new approach involved

an inductive examination of nature by way of observation, of physical

quantities, and experimentation. It helped promote what has come to be known as the

.

Medieval scientific thought prior to Bacon was conducted “…within an

Aristotelian framework with goals very different from Galileo and his successors.”5 In

this system the primary interest was the logical relationships among ideas, while testing

hypotheses by way of experimentation was secondary. Aristotle’s explanations for

natural phenomena were offered in terms of the intelligible essence of objects and their

ultimate purpose. Due to underlying reliance on the rationality of the universe, medieval

science was primarily deductive, beginning with general principles and reasoning to

particular manifestations of those, rather than inductive, beginning with particular

4 Bacon, The New Organon, ed. Fulton H. Anderson (Upper Saddle River, New Jersey: Prentice Hall Publishing Company, 1960), 129. 5 Ian G. Barbour, and Science, Historical and Contemporary Issues (San Francisco: Harper Collins Publishers, 1997), 4. 2 observations and making generalizations based on these. Deductive thought came from the classical idea, prominent with Plato (~ 425-347 BC), that knowledge involves contemplation rather than observation. In other words, the perfect forms of eternal truths have imperfect embodiments in the world in which we live everyday life. Rather than observation of how these imperfect embodiments operate and behave in the changing world, knowledge involves higher contemplation of the perfect forms themselves.6 This is not to say that observation was completely absent from medieval science, as Aristotle offered conclusions that aligned with common experience. In addition, he used detailed observations in order to make various biological classifications.

In the Aristotelian framework humanity was regarded as “…unique and central in both location and importance; the divine was more perfect and comprehensive, separated both geographically and metaphysically from the created order…Everything was in its neatly arranged place in an integrated total plan. It was a law-abiding world, but the laws were moral and not mechanical.”7 This approach primarily regarded mathematics as an independent science with abstractions separate from material bodies that move and interact. Quantities that were the objects of this mathematics were not actually treated as empirical realities. Furthermore, this view identified a more abstract science known as , with objects that were regarded as existing beyond what is subject to treatment by empirical categories. Consider the basic concept of motion as an example.

In an Aristotelian framework motion was regarded as an innate tendency of an object, and was defined as the tendency of an object to find its “natural place” in the universe. In other words, everyday natural objects were considered based on their purpose or inner

6 Ibid., 5. 7 Ibid., 6. 3 cause. These inner causes were said to arise from the “forms” of the objects, so that explanations of everyday processes in terms of the immaterial forms (final causes) were essentially stating “…what has been effected and not an account of factors operative within the .”8

Contrary to the Aristotelian regard for knowledge, Francis Bacon held that the aim of knowledge is not merely or primarily contemplation, but rather practical work in that it is to produce explanations about the way our world works that in turn result in

“…the promotion of happiness and the relief of man’s estate.”9 In his work The

New Organon Bacon shared his six divisions of knowledge where he discussed: 1)

“Divisions of the Sciences,” basically a summary the current scientific knowledge at that time, 2) “Directions Concerning the Interpretation of Nature,” wherein he advocates for the use of inductive reasoning to prepare the human intellect, 3) “A Natural and

Experimental History for the Foundation of Philosophy,” which included a description of some of the elements of what would subsequently develop into the scientific method, 4)

“The Ladder of the Intellect,” which included a discussion of the importance of modeling and typing in science, 5) “Anticipations of the New Philosophy,” which included established scientific proofs, and 6) “The New Philosophy, or Active Science,” or essentially a statement of scientific ways of knowing.10 He proposed “…that the entire work of the understanding be commenced afresh, and the mind itself be from the very outset not left to take its own course, but guided at every step; and the business be done as if by machinery.”11 In this sense he viewed the scientific endeavor primarily as an

8 Bacon, The New Organon, xxiii. 9 Ibid., xxvii. 10 Ibid., 17-29. 11 Ibid., 34. 4 automatic process. While this view emphasized the empirical approach that became key to scientific investigation, it did not really account for the role that theory and creative imagination would subsequently come to play in modern physics.

The birth of modern science, therefore, involved major changes in thought and methodology, along with the associated challenges for the Aristotelian framework for doing science. This was also true for theology, which held to the central significance and importance of humanity, analogously to the Ptolemaic view of as residing at the center of the universe surrounded by the concentric celestial spheres of the heavens. A crucial dilemma for theology that arose during the birth of modern science was “How can theologians make use of the best philosophy and science of their day, as Aquinas made use of Aristotle, and yet avoid distorting the essential Christian message by making an inflexible system that hinders response to new intellectual currents?”12

The revolutionary changes in reasoning brought about by such early scientists as

Copernicus, Bacon, Galileo, Newton and others marked the beginning of further, sometimes rapid, development in scientific thinking. It was further consideration of

Galileo’s insights regarding straight-line motion that led to generalizing these insights for application to motion in general – motion along curved paths. The rigorous quantitative analysis associated with this process of generalization led Isaac Newton and Gottfried

Leibniz (1646-1716) to the discovery of differential calculus. As pointed out above,

Bacon helped lead the recognition that the language of mathematics is the physicist’s tool for modeling reality, rendering conclusions both qualitative and quantitative in nature. It is this same language with which the physicist learns to make predictions about the

12 Barbour, Religion and Science, Historical and Contemporary Issues, 30. 5

behavior of nature. With the advent of the Newtonian view of the world, what we know

today as Newtonian Mechanics, scientists began to believe that an adequate

understanding of at least a part of reality might be within humanity’s reach.

While the work of the early modern physicists was done from curiosity and

empirical motivations to better know how the world works, these individuals also worked

as people with beliefs and values, both religious and non-religious. Some of these were

people specifically with Christian beliefs and values. For the most part these early

scientists saw their scientific endeavors as an aid to their beliefs and not necessarily

inconsistent with them. Bacon warned against mixing science with religion and yet he

still maintained the conviction that scientific conclusions had to be limited by religion.13

From 1618 to 1621 Kepler published the results of his research, using twenty years’

worth of observational data which he obtained from Tycho Brahe (1546-1601), in the

work entitled Epitome of Copernican Astronomy. He approached the study of the

heavens with an empiricist approach while maintaining an ardent religious faith. As an

astronomer he recognized and studied the order and structure observed within the fabric

of physical reality. Concerning the correlation between the period of a planet’s orbit and

the planet’s mean distance from the sun Kepler admitted to being “…carried away by

unutterable rapture at the divine spectacle of heavenly harmony.”14 That Kepler viewed

his empirical research as an aid to a more complete understanding of God and the

Scriptures can be inferred from his statement that “My wish is that I may perceive the

13 , Science and Religion (, United Kingdom: Cambridge University Press, 2014), 76. 14 John Brooke, “Science and Religion: Lessons from History?” Science 282, no. 5396 (Dec. 11, 1998), http://science.sciencemag.org/content/282/5396/1985. 6

God whom I find everywhere in the external world in like manner within me.”15

Subsequently, Galileo’s Dialogue on the Two Great World appeared in 1632.16

Like Kepler, Galileo emphasized empirical investigation of the world around us while

maintaining his religious faith. He believed the physical universe was a manifestation of

God’s work, and held that empirical knowledge gained through a study of the physical

universe aligned with the true sense of the Scriptures and the intention of the inspired

writers. He wrote:

...in discussions of physical problems we ought to begin not from the authority of scriptural passages but from sense-experiences and necessary demonstrations; for the holy Bible and the phenomena of nature proceed alike from the divine Word the former as the dictate of the Holy Ghost and the latter as the observant executrix of God’s command. It is necessary for the Bible, in order to be accommodated to the understanding of every man, to speak many things which appear to differ from the absolute truth so far as the bare meaning of the words is concerned. But Nature, on the other hand, is inexorable and immutable; she never transgresses the laws imposed upon her, or cares a whit whether her abstruse reasons and methods of operation are understandable to men. For that reason it appears that nothing physical which sense- experience sets before our eyes, or which necessary demonstrations prove to us, ought to be called in question (much less condemned) upon the testimony of Biblical passages which may have some different meaning beneath their words.17

The very notion of law, which Galileo referred to as being imposed upon nature, and that

has become one of the most fundamental concepts of the natural sciences, has been

regarded by scientists throughout the years as “…a somewhat mysterious concept that

invites a philosophical or even religious interpretation.”18 Bacon, who advocated a

methodical, empirical approach to scientific investigation, also claimed that next to the

word of God physics is a preventative against superstition as well as a nourishment to

15 Will and Ariel Durant, The Story of Civilization: The of Reason Begins (New York, New York: MJF Books, 1961), 600. 16 George O. Abell, Exploration of The Universe (New York: Holt, Rinehart, and Winston, 1974), 53. 17 Galileo Galilei, “Letter to the Grand Duchess Christina of Tuscany, 1615,” Modern History Sourcebook, last modified January 21, 2020, https://sourcebooks.fordham.edu/mod/galileo-tuscany.asp. 18 Joshua Moritz, Science and Religion: Beyond Warfare and Toward Understanding (Winona, Minnesota: Anselm Academic, 2016), 43. 7

faith. He characterized the relationship between physics and faith by saying that physics,

(which he referred to as natural philosophy), “…is rightly given to religion as her most

faithful handmaid, since the one displays the will of God, the other his power.”19

In 1687, Isaac Newton published his Philosophiae Naturalis Principia

Mathematica, marking the crowning point for the of mechanics. Newton too

believed that the universe was a manifestation of the work of an intelligent, all-powerful

Being. While he regarded mathematics as a tool for modeling observed phenomena in

nature, Newton did not view mathematical laws as evidencing a purely autonomous

universe. He understood that “This most elegant system of the sun, planets, and comets

could not have arisen without the design and dominion of an intelligent and powerful

being.”20 For Newton, the world, with laws able to be expressed mathematically,

evidenced God’s purposes.21 Rather than a part of the physical universe or a force within

it, God was, in Newton’s view, separate from it. He perceived God to be supreme,

eternal, and absolutely perfect being. In his words:

And from true lordship it follows that the true God is living, intelligent, and powerful; from the other perfections that he is supreme, or supremely perfect. He is eternal and infinite, omnipotent and omniscient, that is, he endures from to eternity, and he is from infinity to infinity; he rules all things, and he knows all things that happen or can happen. He is not eternity and infinity, but eternal and infinite; he is not or space, but he endures and is present.22

In opposition to the deistic view, Newton sought for evidence for God’s involvement in

the world. It has been argued that his voluntarist theology placed Newton in a quandary

because it allowed that events in nature could be explained by way of both the mechanics

19 Bacon, The New Organon, 88-89. 20 Isaac Newton, The Principia: Mathematical Principles of Natural Philosophy, trans. I. Bernard Cohen and Anne Whitman (Berkeley, Los Angeles, London: University of California Press, 1999), 940. 21 Barbour, Religion and Science: Historical and Contemporary Issues, 18. 22 Newton, The Principia: Mathematical Principles of Natural Philosophy, 941. 8 of the world as well as the will of God. The resulting difficulty lay in being able to determine which events in nature were due to God’s will and which were due to the operations of the laws of nature.23

It is important to realize that from the beginning the birth of modern science involved an empirical approach to understanding and modeling reality. These early physicists held the view that scientific knowledge of the world was not inconsistent with or contradictory to their beliefs and values. Furthermore, scientists of religious faith tended toward the view that knowledge obtained by way of the scientific method could inform one’s faith. Actually the concept of faith lies at the core of the scientific endeavor itself, for “Without the belief that it is possible to grasp the reality with our theoretical constructions, without the belief in the inner harmony of our world, there could be no science.”24 Accordingly, these early physicists dealt with the challenges that came along with newer, more contemporary scientific thought. While their beliefs were not the motivation for these first modern physicists to “do science,” neither did they generally hold to the notion that their beliefs and values were rendered irrelevant by the process of scientific reasoning. In that they adhered to the principle that knowledge of the world around us comes from that world by way of observation and experimentation, it can be argued that the early physicists were among the first “bottom-up thinkers,” along with others who preceded them with their efforts to adopt observation as the starting point for doing science. They sought to know the world by way of empirical reasoning, yet through the maturity of knowledge gained in this way many of them also endeavored to come to a more complete, or more adequate knowledge of God’s relation to the world.

23 Brooke, Science and Religion, 196. 24 Einstein, Infeld, The Evolution of Physics: From Early Concepts to Relativity and Quanta, 296. 9

Viewing Physical Reality as a Great Machine

Newtonian mechanics far excelled all other explanations which came before, for

at least a couple of reasons. First, it was based solidly on the results of experiments

performed by Galileo and subsequently by others. This was a marked improvement over

the conjecture and speculation that formed the base of earlier theories. ,

Newtonian mechanics did not only concern itself with the conditions prevailing at the

surface of Earth (boundary conditions). Therefore, it was able to provide explanations

and make predictions for observational astronomy. Newtonian mechanics was seen to

apply to the known universe, not merely Earth.25 This was a great leap for scientific

theorizing, and one that no doubt helped to establish the notion that the Newtonian view

was definitive regarding the essence and behavior of the natural world.

Moreover, as with any area of scientific thought Newtonian mechanics made

some fundamental assumptions, among which were that 1) space provided an absolute

background for all the observed phenomena in nature, and 2) all natural events could be

measured against the constant procession of time. For Newton, space and time were

absolutes, not variables that could potentially affect experimental results. He viewed

space and time as constituent parts of the grand laboratory in which natural phenomena

are observed and measured, rather than natural phenomena subject to change and

alteration.

At the time of its establishment, Newtonian mechanics offered humanity a

potential power unprecedented in history – the theoretical ability, assuming one had

25 James Jeans, Physics and Philosophy (New York, New York: Dover Publications, 1981), 108-109; and , Out of My Later Years (New York, New York: Citadel Press, 1984), 222. 10 enough information at hand, to calculate to the finest detail the state of the world as well as the rate at which this state will change. Proceeding backwards one should be able to trace the chain of causation back to the origin of the system. Moving forward one should be able to determine each successive change indefinitely, so that an accurate picture of the could be established. Such optimism led the French mathematician Pierre-

Simon de Laplace (1749-1827), in 1812, to say of the fictional, infinitely capable, and industrious mathematician with sufficient information regarding his present state that

“Nothing would be uncertain for him; the future as well as the would be present to his eyes.”26 What was important for this view was not the non- of such a mathematician, but rather the determinate nature of the universe. According to this view of the world, all information regarding the system is inherently part of its original configuration so that its future unfolds according to applicable natural laws.

The point has been made that “In 1704 a contributor to a French learned journal observed that a new style of scientific observation had become the rage. One heard of nothing else but mechanistic physics.”27 This view of the world, sometimes referred to as

“mechanistic determinism,” did not burst onto the scene as fully developed concurrently with Newtonian mechanics. Rather it derived its existence from the same fundamental assumptions on which Newton based his system of thought, and continued to grow and develop into a worldview as it was proven possible to extend Newtonian mechanics to account for a wider range of natural phenomena – a process generally known as “classical mechanics.” The English physicist James Jeans (1877-1946) explained mechanistic determinism as a view where:

26 Ibid., 111. 27 Brooke, Science and Religion, 158. 11

It follows that the changes of the world at any instant depend only on the state of the world at that instant, the state being defined by the positions and velocities of the particles; changes in position are determined by the velocities, and changes in the velocities by the forces, which in turn are determined by the positions.28

The hope offered by classical mechanics was that in time not only a wider range of natural phenomena, but all natural phenomena, would be explained in terms of the mechanical principles given by Newton’s approach. With this confidence in physics’ ability to eventually explain all, mechanistic determinism became firmly entrenched as the predominant scientific worldview. In the words of Einstein and Infeld:

The great achievements of mechanics in all its branches, its striking success in the development of astronomy, the application of its ideas to problems apparently different and non-mechanical in character, all these things contributed to the belief that it is possible to describe all natural phenomena in terms of simple forces between unalterable objects. Throughout the two following Galileo’s time such an endeavor, conscious or unconscious, is apparent in nearly all scientific creation.29

The successes of the first modern physicists were so great that the scientific method seemed to take on a life of its own as its approach to obtaining knowledge of the world around us was adopted by other scholarly disciplines. During the seventeenth, eighteenth, and nineteenth centuries it was as if the scientific method adapted itself to scholarly pursuit in nearly every field of human intellectual study. This “unity of knowledge” approach, which is the belief that truth in all fields of human inquiry is obtainable through one methodology, was a natural outgrowth of the success of the scientific method. The resulting proliferation of scientific methodology, along with its success in providing explanations, also contributed to “,” which is the notion that all reality can eventually be reduced to the laws and principles governing material existence. In other words, there was a growing confidence that 1) Newtonian physics can

28 Ibid., 109. 29 Einstein, Infeld, The Evolution of Physics: From Early Concepts to Relativity and Quanta, 42-43.

12 eventually explain all natural phenomena, 2) all things that exist are natural, and therefore

3) Newtonian physics can eventually explain all things.

This tendency toward focusing on the material was fundamentally a departure from the worldviews of the earliest modern physicists who, for the most part, gave due regard to the spiritual. It was not, and is not, uncommon for scientific research to yield results that generate changes in philosophical perspectives relating to questions and problems that extend beyond the boundaries of science. The body of knowledge accumulating in the field of physics provided the material for a mechanical, deterministic view of the world to develop and thrive. Though not the only worldview, mechanistic determinism became a predominant worldview grounded in scientific thought and principles. It continued to be intimately tied to the scientific method even though its particular concerns sometimes exceeded the boundaries of scientific research.

This view of the world as machine, then, did not regard the question of divine interaction with the world as particularly relevant or even valid. The new implied question which arose was “How can there be room for divine interaction in a world that runs with clockwork precision?” As more of the “machinery” of reality became comprehensible, adherents of this philosophical perspective became more confident in humanity’s ability to answer the “ultimate” questions without inclusion of the spiritual.

For the mechanistic determinist even what people regarded as the spiritual was a realm of phenomena that could eventually be defined and explained in terms of matter and energy, and applicable physical laws. It seems ironic that this purely materialistic view came to be known as the Newtonian view, since Isaac Newton himself did not maintain a purely materialistic view of reality.

13

Subsequent rethinking of Newton’s fundamental assumptions regarding space and time helped bring about revolutions in twentieth-century physics. Change and development characterize knowledge gained through the scientific method, for “There are no eternal theories in science. It always happens that some of the facts predicted by a theory are disproved by experiment. Every theory has its period of gradual development and triumph, after which it may experience a rapid decline.”30 Not only is this true of individual theories; it is also true of entire systems of thought. The classical mechanical view entered the twentieth century bent on explaining all encountered phenomena in mechanical terms. However, the twentieth century brought with it phenomena (electrical, magnetic, and optical phenomena among them) manifesting behaviors that were problematic to explain in a purely mechanical view of the world with its reliance on particles, their behaviors, and positions with respect to time. As a result, new approaches were taken toward understanding and explaining these phenomena. These new approaches often yielded insights that in turn made it possible to provide explanations of the new phenomena, as well as older phenomena, without having to rely on the classical mechanical view. This eventually led to two great revolutions in the twentieth century that altered physics’ view of physical reality.

Losing the Mechanical View of Physical Reality

We can consider these revolutions in modern physics by first considering problematic behaviors manifested by electrical phenomena. The simple theory offered up to explain observed behaviors during experiments with electrical phenomena proposed

30 Einstein, Infeld, The Evolution of Physics: From Early Concepts to Relativity and Quanta, 75. 14 the existence of two electric fluids. These fluids were regarded as being similar to

(which originally was regarded as a substance, but a substance without ) in that the amount can change, or flow, from object to object, while the total amount in an isolated system remains constant. The two electric fluids, called charges, were designated as positive (+) and negative (-), with the understanding that like charges repel and unlike charges attract. Here, then, was a mechanical notion of force. Materials allowing freedom of movement to electric charges were termed “conductors,” while those tending to inhibit movement of charges were called “insulators.”

The fluids theory of electrical phenomena is indicative of the classical approach of explaining natural phenomena in terms of substances and the forces at work between them. Newton’s universal law of gravitational force states that the force of is directly proportional to the sum of the two involved and inversely proportional to the square of the distance between them. A century after Newton came upon this law,

Charles A. de Coulomb (1736-1806) discovered a similar dependence on distance in the laws governing attraction and repulsion of electric charges. However, while gravitation is strictly attractive and manifests in the presence of objects of mass, electric forces were found to be attractive or repulsive, and only existing in the presence of charge. The same question arose with as did with heat. Namely, do electric charges possess mass? Just as a heated object was not observed to register differently on the scale when it was allowed to cool, or when it was heated more, electrical fluids appeared to “carry no weight” – charged objects did not change the scale.

Furthermore, as scientists made a distinction between temperature and heat when examining heat phenomena - Temperature is the average of a collection of

15 molecules of a given substance, while heat is the total internal energy of a collection of molecules undergoing transfer from one substance to another - they also made a distinction between “electric potential” and “electric charge” within the realm of electrical phenomena. Without going into detail, it can be shown that two spheres of different radii can have the same electric charge while possessing different electric potentials. Electric potential is directly related to the density of the charges and is a measure of the likelihood that electric charges will be forced to move from one point to another. Physics texts occasionally use the analogy of a water system where the pump represents electric potential (voltage) and the flow of water represents electric current.

In a manner similar to electric phenomena, the mechanical view was also applied to magnetic phenomena. Simple experimentation with two bar magnets showed the existence of the familiar attraction and repulsion forces between magnets. We say that a magnet possesses two “poles,” designated north and south, and that unlike poles attract while like poles repel. However, something unexpected occurs when one of the bar magnets is broken. Instead of ending up with one pole (like one lone electric charge), there are still two poles – essentially a smaller bar magnet. Magnets are constituted such that magnetic dipoles are inherent in the material of the magnet itself, so that continual breaking just produces smaller and smaller magnets all the way down to the molecular level. As with gravitation, and electrical phenomena, magnetic forces too were found to have a dependence on distance that is inversely proportional to the square of the distance between the magnetic poles.

Knowledge that electrical and magnetic phenomena are not two aspects of the same thing was evidenced by the fact that a stationary electric charge near a magnet

16

experiences no force, nor does the magnet. However, during a lecture demonstration in

the 1820s, the Danish physicist Hans Christian Oersted (1777-1851) discovered that a

compass needle was deflected by a current-carrying wire.31 In other words, if an electric

charge is at rest near a magnet, there are no mutual forces, but set the charge in motion

(or the magnet, or both) and suddenly there exist mutual forces acting on each. If the

charge exerts a force on the magnet, the magnet also exerts a force on the charge.

Furthermore, the mutual forces created by way of the relative motion between the magnet

and charge do not act along the centerline joining the two. Rather they act along a line

perpendicular to the centerline.

The classical view maintained that all natural phenomena could eventually be

explained in terms of particles and the forces acting between them. In addition, this

mechanical view of nature regarded all such forces as acting along the centerline between

the objects, and as being dependent solely on the distance between the objects. Then a

“force of a different color” appeared which depends on the relative velocity of the

objects, and does not even exist if the objects are at rest with respect to each other. To

confound matters more, this new type of force does not even act along the objects’

centerline. The force(s) arising due to moving charges and/or magnets confounded the

classical view but did not deal it a deathblow. Explanations for these new forces can be

framed, though with some complexity, in terms of mechanics. Subsequent phenomena

were much more problematic and demonstrated the necessity to relinquish a hold on the

view of nature as strictly mechanistic. However, it seems that the new electrical and

magnetic forces were harbingers of what was yet to come.

31 Raymond A. Serway, Physics for Scientists and Engineers (Philadelphia: Saunders College Publishing, 2000), 905. 17

Another realm worth considering here is that of light. Galileo posed the question

as to whether or not light possesses velocity - Does it require a finite amount of time for

light to travel from one point to another, or is it an instantaneous ? In his

Dialogue Concerning ,32 he wrote of a conversation between a teacher

and his pupils regarding the of light:

Sagredo: But of what kind and how great must we consider this to be? Is it instantaneous or momentary or does it like other require time? Can we not decide this by experiment? Simplicio: Everyday experience shows that the propagation of light is instantaneous; for when we see a piece of artillery fired, at great distance, the flash reaches our eyes without lapse of time; but the sound reaches the ear only after a noticeable interval. Sagredo: Well, Simplicio, the only thing I am able to infer from this familiar bit of experience is that sound, in reaching our ear, travels more slowly than light; it does not inform me whether the coming of the light is instantaneous or whether, although extremely rapid, it still occupies time. An observation of this kind tells us nothing more than one in which it is claimed that "As soon as the sun reaches the horizon its light reaches our eyes "; but who will assure me that these rays had not reached this limit earlier than they reached our vision? Salviati: The small conclusiveness of these and other similar observations once led me to devise a method by which one might accurately ascertain whether illumination, i.e., the propagation of light, is really instantaneous.

The account proceeds with a description of Salviati’s proposed experimental

configuration. We will not concern ourselves with the details here. Galileo was unable,

with the resources available to him, to conceive of an experimental set up that would

render a value for light’s velocity. However, with time the velocity of light has been

determined repeatedly, with ever-increasing accuracy. Today we know, through

experimental research, that light possesses a finite velocity of approximately 3 x 108 m/s

32 Galileo Galilei, Two New Sciences, trans. Henry Crew, Alfonso de Salvio (New York: The Macmillan Company, 1914), 42, https://www.stmarys- ca.edu/sites/default/files/attachments/files/Dialogues_Concerning_Two_New_Sciences.pdf. 18

in a vacuum.33 Though the question of the velocity of light was somewhat puzzling to

physicists, it was the question of the nature of light that had greater implications for an

understanding of the structure of reality.

In the early days of the classical view two main conceptions of light were vying

for general acceptance – the corpuscular theory of light, advocated by Newton, and the

wave theory of light, the proponent of which was Christian Huygens (1629-1697), a

contemporary of Newton. The corpuscular conception of light regarded light as

essentially a collection of particles, or corpuscles. This was in keeping with the

observation that light travels with rectilinear (straight-line) motion. In fact, it was light’s

rectilinear motion that cemented Newton’s belief in the corpuscular theory, even though

he understood the benefits of Huygens’ wave conception.34 However, the assumption of

the straight-line propagation of light had its limitations. Generally, this is a valid

assumption for light travelling in a vacuum, but not for other materials through which

light may travel. When light enters a material medium it has a tendency to alter its path,

a phenomenon known as refraction. This characteristic of light was more difficult for the

corpuscular theory to handle, but Newton did offer the explanation that the inter-particle

forces between the corpuscles of light and the atoms of the material attracted the

corpuscles toward the normal (a line perpendicular to the surface of a material). Newton

had to assume that light traveled more quickly in a material medium than it did in a

vacuum, an assumption later proved false. In a similar fashion Newton explained that

light could be separated into its constituent particles by a prism, producing a range of

33 Raymond A. Serway, Physics for Scientists and Engineers, 1108. 34 Paul A. Tipler, Physics for Scientists and Engineers (New York, New York: Worth Publishers, 1991), 974. 19 colors.35 This complicated matters because it not only required light to consist of particles, but myriad of types of particles, each type being associated with a different color.

Huygens conceived of light as a wave in nature. A wave can be defined as "a disturbance or oscillation propagated with a definite velocity from point to point in a medium.” Thinking of a wave as a series of valleys and crests moving in one direction, the period (T) is the time required for the wave to travel one wavelength (the distance from crest to crest or valley to valley). The frequency () is the number of wavelengths passing a point each second. Some common behaviors of wave phenomena include:

1. reflection: bouncing off an object in its path 2. diffraction: bending around an object in its path 3. refraction: changing direction on entering a different medium 4. interference: arriving at a point at the same time as another wave 5. energy possession/translation

Basically, a wave travels in one of two ways. It can move such that the direction of motion is the same as the direction of its oscillation (longitudinal wave), or it can move such that the direction of motion is perpendicular to the direction of its oscillation

(transverse wave).

Time will not afford the luxury of a point-by-point comparison of the corpuscular and wave theories of light. Most of the observed behaviors of light were capable of explanation by both up until the middle of the nineteenth century. By then the scientific community opted for the wave theory because of its ability to go further to explain a behavior of light not observed in Newton’s day – diffraction. Diffraction, as defined above, is the ability of a wave to bend around an object in its path. The corpuscular

35 Einstein, Infeld, The Evolution of Physics: From Early Concepts to Relativity and Quanta, 95-97. 20 theory of light could not account for this behavior among particles. Particles of matter simply do not behave in this way. However, waves do, and this can very easily be demonstrated in the lab using water as the medium. Not only did experimental data regarding diffraction confirm the wave nature of light, they confirmed light to be a transverse wave.

Still, the traditional way of thinking regarded light waves as essentially mechanical, for the propagation of a wave for some distance required a medium, some material or substance, which interacts with the original disturbance or oscillation. It is the presence of the medium that makes wave propagation possible from the mechanical point of view. Therefore, great and various attempts were made to describe the mechanical nature of the medium that allowed the propagation of light. This medium, which came to be known as the “ether,” was initially regarded as the ultimate or absolute for describing events in the universe. If one recalls that Newton regarded space to be an absolute background against which all events in nature take place, then the existence of the ether implied that there was really no such thing as “free space” (empty space).

Serious problems began to arise in the course of trying to identify the exact nature of the ether. From mechanics it was known that celestial objects experience no resistance in their trajectories. The influences they know come from other objects by way of gravitation, not from space itself. Yet there must be some type of interaction between the ether and light waves if the waves are to travel. In other words, the ether needed to be inert in its encounter with mechanical objects such as planets and , but not in its encounter with mechanical phenomena such as light waves. For the first time in its

21 history, the mechanical view encountered questions that had the potential of completely revising fundamental assumptions upon which the mechanical notion of nature was based. Subsequently, changing assumptions regarding space and time rendered a definitive verdict regarding the ether, as well as brought about changes in the assumptions concerning the notion of determinacy within nature.

22

CHAPTER 2

MODERN PHYSICS’ VIEW OF REALITY – NEW CONTEXTS FOR OLD

QUESTIONS

Looking back on the changing assumptions concerning various processes and systems in physical reality that led to rethinking viewing the world as a great machine, it can be argued that these were harbingers of more changes, greater changes, to come.

Adopting new ways of thinking about phenomena not easily characterized by way of a strict mechanical view of the world, resulted in continued revision of modern physics’ description of physical reality. A review of some of these new ways of thinking will help to show how they brought about the two great revolutions in modern physics during the twentieth century that were mentioned earlier on.

The concept of “field” is one example of modern physics’ new way of regarding the world around us. A new force was discovered which refused to conform to the traditional mechanical view of the world as a collection of particles and the distance- dependent forces acting between them. In particular, this is the force experienced by a magnetic pole due to a moving electric charge (current). This force is velocity dependent and is directed perpendicular to the line connecting the charge and magnetic pole.

Pictorially we can represent this force as a line or curve with associated arrows indicating the direction in which the force acts. For a straight current-carrying wire, the lines of force due to the moving charges are represented as circular paths around the wire. To illustrate this, with your left hand point your thumb in the direction of the moving charge.

Your fingers will curve around the wire in the direction of the lines of force. The arrows

23 indicate the direction in which the force acts, while the density of the lines is directly related to the strength of the force. As this force is magnetic in nature, these lines of force represent the “magnetic field” associated with moving charges. The field representation yields information about the force acting on a magnetic pole at any given point in space.

For a current-carrying coil of wire (pictured as looking somewhat like a “slinky”) the associated magnetic field runs from positive to negative such that the magnetic field representation looks much like the magnetic field representation of a bar magnet. In fact, ignoring the coil and magnet results in representations of two identical magnetic fields.

Without the concept of field, it would be difficult, if not impossible, to perceive and describe this similarity. Experiment bears out the similarity, for the behavior of a current-carrying coil and a bar magnet are exactly the same. This is explainable by the fact that the fields are responsible for these behaviors, and the fields are of exactly the same character irrespective of their sources.

The concept of field is not limited to the interaction between electric charges and magnetic poles. It has actually proven to be useful for describing various aspects of physical reality, including gravitation. An electric field exists by virtue of the existence of electric charge. Pictorially this is represented by a small sphere, which indicates the charge (a point source), and radiating lines perpendicular to the sphere’s surface, which indicate the lines of electric force, or “field lines.” For a positive charge the field lines radiate outward in all directions and for a negative charge they radiate inward from all directions. A would be diagramed much like the field of a negative electric charge – the sphere in this case would represent a body of matter (such as a

24 planet) and the inward-radiating lines the gravitational field, which always acts toward the sphere’s center.

Using the concept of field to describe electrical-magnetic phenomena leads to the conclusion that an electrostatic field (unchanging electric field) exerts no influence on a magneto-static field (unchanging magnetic field) and vice versa. However, when the electric field changes in some way, such as occurs when it is set in motion, it brings about a magnetic field. In other words, a changing electric field induces a magnetic field.

Furthermore, if the frequency of the change, or the magnitude of the change, is increased, then the strength of the associated magnetic field is also increased. Early in the nineteenth century Michael Faraday (1791-1867) and Joseph Henry (1797-1878) independently verified by experiment that a changing magnetic field produces an electric field.36 In a similar manner to the reverse case above, increasing either the strength or frequency of the change of the original magnetic field causes the associated electric field to increase in strength.

There is another result of utilizing the field concept to explain phenomena formerly modeled by particle-particle interactions, and that is regarding the field as a source of energy. There is an evident symmetry in the phenomena discovered by Oersted and Faraday, which can be explained as energy conservation. Energy is a constituent of the fields, whatever type they might be. Einstein and Infeld wrote: “The attribution of energy to the field is one step further in the development in which the field concept was stressed more and more, and the concepts of substances, so essential to the mechanical view, were more and more suppressed.”37 These field phenomena led James Clerk

36 Tipler, Physics for Scientists and Engineers, 782. 37 Einstein, Infeld, The Evolution of Physics: From Early Concepts to Relativity and Quanta, 142. 25

Maxwell (1831-1879) to formulate the laws governing the structure and behavior of electromagnetic fields.38 This occurred late in the nineteenth century, and was regarded by Einstein to be the most important in physics since the works of Newton.39

Maxwell’s equations define the electromagnetic field as something real irrespective of any test charge or pole. In other words, Maxwell’s theory states that a changing electric field produces a magnetic field and a changing magnetic field produces an electric field, even though there may exist no respective magnetic poles or electric charges for detecting these fields.

Maxwell’s theory yielded consequences crucial for comprehending natural phenomena without appeal to a mechanistic view. One of these consequences was the ability to account for wave propagation in free space. The field represents energy and a changing electric field produces a changing magnetic field, which produces a changing electric field, and so on. A source of a changing electric field, such as an oscillating electric charge, will produce an associated electromagnetic wave by virtue of the changing electromagnetic field that results. This wave, in turn, propagates through space with a definite velocity as alternating electric and magnetic fields, without the need for a medium made of particles. Since the electric fields and magnetic fields are orthogonal, or perpendicular, to each other, the electromagnetic wave propagates as a transverse wave, which agrees with earlier conclusions about light. A further consequence of this theory, which is supported by experimental data, is that the electromagnetic wave travels through space with the same velocity as light. In other words, light is an electromagnetic wave.

Clearly Maxwell’s field theory established a link between electromagnetic and optical

38 Tipler, Physics for Scientists and Engineers, 782. 39 Einstein, Infeld, The Evolution of Physics: From Early Concepts to Relativity and Quanta, 143. 26 phenomena, realms previously thought to be unrelated. This is perhaps one reason why

Einstein and Infeld stated that: “The theoretical discovery of an electromagnetic wave spreading with the speed of light is one of the greatest achievements in the .”40

The field concept began as a way to conceptualize natural phenomena from the mechanical perspective. However, the language of field theory made it clear that the field itself, rather than the particles with which it is associated (i.e. electric charges, bodies of matter in space, etc.), is a reality giving rise to forces in nature. Field began to eclipse particle as aspects of natural phenomena came to be explained without reference to particles and the distance-dependent forces acting between them. The concept of field came to be instrumental in explaining phenomena that were previously explained using the mechanical view of nature. This is not to say that the concept of field definitively abolished the mechanical perspective. However, it did result in bringing greater focus on the limitations and shortcomings of this view. According to Einstein it rendered a wider view that still allowed for the retention of some of the old conceptualizations.41 For example, Maxwell’s theory maintains the idea of electric charge, but as a source of the field rather than a particle or substance. Still the advent of the notion of field demonstrated the changing climate for modern scientific investigation of the world, a climate fitted for revolutionary thinking that would bring about the replacement of the mechanical view.

The mechanical view’s reliance on particles and their behaviors to explain natural phenomena required a way to determine and express the particles’ positions with respect

40 Einstein, Infeld, The Evolution of Physics: From Early Concepts to Relativity and Quanta, 151. 41 Ibid., 152. 27 to time. This was accomplished with the aid of “frames of reference” or “coordinate systems.” A is essentially a fixed, rigid background or grid serving as the arena in which particles operate. The coordinate system provides a way to quantify position, velocity, and acceleration, of the associated particles. One result of the mechanical view is the prediction that mechanical laws valid in one coordinate system are valid in all coordinate systems moving uniformly (with constant velocity) with respect to the original (such coordinate systems are termed “inertial” coordinate systems).

This is known as the “Galilean relativity principle.”42

The Loss of the Constancy of Space and Time in Physics

Albert Einstein (1879-1955) worked to answer questions about the very nature of space and time through his special . Accordingly, he made two assumptions regarding the nature of the ether mentioned earlier. First he assumed that moving bodies carry the ether along with them. For this to be true the velocity of light emitted from a moving light source would differ from the velocity of light emitted from a stationary source – both measured by an external observer at rest. However, experiment showed the velocity of light to be independent of the state of motion of the light source.

The implication, from the perspective of a mechanical view of the universe, is that the ether is fixed (in accordance with Newton’s conception of space). Therefore, the second assumption was that the ether defines a unique coordinate system through which moving bodies travel. The truth of this assumption means that the velocity of light emitted from a moving light source, as measured by someone moving with the source, must differ from

42 Ibid., 153. 28 the value obtained by an observer at rest, invalidating the Galilean relativity principle.

However, experiment again had a different conclusion, namely that both observers register the same value for the velocity of light. In 1887 two American scientists, Albert

Michelson (1852-1931) and Edward Morley (1838-1923), performed a crucial experiment with regard to the ether as a fixed universal frame of reference – the famous

Michelson-Morley experiment. Their experimental results indicated that there exists no absolute frame of reference in the universe such as would be rendered by the ether.

Therefore, Einstein said that there is no frame of reference that is absolutely at rest, or, stated another way, there is no coordinate system that allows the determination of absolute motion. All motion in the universe is relative motion. The mechanical concept of ether had proved useless as an aid to explain electromagnetic phenomena in moving coordinate systems. , with earlier help from Michelson and Morley, dealt the concept of ether a deathblow.

As already stated, the natures of space and time were subject to review by

Einstein’s special theory of relativity, which began with the following assumptions:

1. The velocity of light in vacuo is the same in all coordinate systems moving uniformly, relative to each other. 2. All laws of nature are the same in all coordinate systems moving uniformly, relative to each other.43

This means that two observers, one inside a moving room containing a light source in the center, and the other outside of the moving room and at rest, will measure a beam of light emitted from the source as traveling with exactly the same velocity. However, the internal observer will see the light beam strike the walls of the room simultaneously, whereas the outside observer will see the “approaching wall” encounter the light beam

43 Ibid., 176. 29 slightly before the “escaping wall” does. The implication, then, is that time does not transform in the classical sense (the sense in which Newton thought of time as progressing at a universally constant rate) between two reference frames moving relative to each other. Special relativity says that two events that are simultaneous in one frame of reference are not necessarily simultaneous in another. In other words, time in a moving frame of reference is not the same as time in a stationary frame of reference. In particular, in moving frames run slower than those in frames at rest.

Furthermore, the assumptions of special relativity lead to conclusions regarding spatial, as well as temporal, dimensions. In particular, Einstein was able to show by way of his theory that along the axis of motion contracts in a moving frame of reference. This contraction is directly proportional to the reference frame’s velocity so that as the velocity of the moving frame increases, so does the contraction. In other words, material objects contract in the direction of motion in moving frames of reference.

Stated another way, space in a moving frame of reference is not the same as space in a stationary frame of reference.

It is important to note here that the translation of space and time from one frame of reference, at rest, to another moving with respect to the first, behaves much like

Newton’s conception of space and time when the relative velocity of the two frames is small compared to the velocity of light. That is, for velocities that people encounter in everyday life, the difference in spatial and temporal dimensions between two frames of reference moving relative to each other is negligible. However, when the relative velocity between two reference frames approaches that of light, the change in space / time from one frame to another is extremely noticeable. The view of special relativity is

30 a more accurate description of reality and yet the classical view is a limiting case of this more accurate view of nature.

One other implication brought about by special relativity has to do with the relationship between matter and energy. The classical view introduced these as two types of “substances” – one with mass, and one without. These substances were observed to interact with each other, and it was seen that certain conservation laws were observed.

That is, either substance may undergo certain transformations or translations during its varied interactions, but the total amount of substance remained essentially the same. So, per the classical view there was one conservation law for matter and one for energy.

However, the assertions of special relativity introduced a new conceptualization with respect to the interaction of matter and energy, namely that all energy is resistant to change in the state of motion, and in that sense behaves like matter. In fact, Einstein discovered that matter and energy are essentially the same – two manifestations of one more fundamental reality. Therefore, the old classical conservation laws were replaced by a single law of conservation of matter and energy.

First, Einstein showed by way of his theory of special relativity that physical laws are valid in inertial coordinate systems (coordinate systems moving uniformly - with constant velocity - with respect to each other) without reliance upon the mechanical view.

Second, he generalized his findings by formulating the validity of physical laws in all coordinate systems. This generalization has come to be known as . That is, the translation of space and time between moving frames of reference is valid for all motion. Special relativity defines how space and time translate between two relatively moving frames when that motion is uniform, whereas general relativity defines the

31 translation of space and time between two relatively moving frames when the motion is accelerated.

The view of reality as conceptualized by the special and general theories of relativity constitutes the first revolution in modern physics. The results of Einstein’s work revealed space and time as sense experiences, rather than the absolute background against which sense experiences are perceived. Rather than viewing nature as a great spacious laboratory of three spatial dimensions, in which anything can be measured in time, relativity rendered the universe as a four-dimensional (three spatial dimensions and one temporal dimension) space-time continuum. With the advent of relativity, space and time lost absoluteness and constancy, and became subject to alteration depending on the observer’s perspective, as well as being seen as interwoven into the single fabric of physical reality. In addition, this space-time continuum is curved, so that a beam of light traveling in a gravitational field will “bend.” One of the experimental confirmations of the curvature of space-time is this observed “bending” of light rays by the sun. When the sun is aligned such that the light from some distant passes near its edge, such light is diffracted - its path of travel is altered. It has been shown that light itself has no mass, but rather consists of pure energy. Still light traveling through space is curved by the presence of a mass in space. The result of this experimental verification is that the light source, the star, appears to be positioned in a different location in space than it actually is.

This phenomenon can be verified by making observations of the stars in that same region of deep space at other of the , when the sun is not “in the way” of the distant star’s light.

32

The Loss of Determinacy in the Fabric of Space-Time

A second revolution in physics’ view of reality came about as a result of the

classical mechanical view’s inability to account for the stability of the atom. According

to the classical mechanical view of an atom, the negatively charged electrons should fall

into the positively charged nucleus (because their attraction should eventually overcome

the state of orbital motion), otherwise the electrons would have to remain in perpetual

motion, which is not allowed according to the energy conservation laws of classical

mechanics. (1885-1962) found that the permanence of the atom, rather than

its existence, called for the need for a revision of the classical view.44 This eventually led

to the discovery of energy quantization with respect to light phenomena, as well as an

inherent wave-particle duality in nature, otherwise known as the principle of

complementarity. 45 Furthermore, it was found that this complementarity in nature is not

restricted to light, but is also observed in the fundamental subatomic constituents of the

atom. It is interesting to note that the argument has been made that with regard to

articulating this complementarity in nature Bohr was influenced by considerations from a

variety of other disciplines outside of physics, including physiology (Bohr’s father held

the chair in physiology at the University of Copenhagen), psychology, and even

.46 The study of physical reality in light of its innate quantum

behavior came to be known as “quantum mechanics.” The term “quantum mechanics”

was not intended to indicate a mechanical view of nature, but rather was an expression

44 Niels Bohr, “The Structure of the Atom,” Nobel Lecture, Royal Swedish Academy of Sciences, Stockholm, Sweden, Dec. 11, 1922, 11, https://www.nobelprize.org/uploads/2018/06/bohr-lecture.pdf. 45 Ibid., 14; and , Physics and Philosophy: The Revolution in Modern Science (Amherst, New York: Prometheus Books, 1999), 43. 46 Brooke, Science and Religion, 453. 33 for a new arena in which physical reality could be investigated and in which the rules long established for explaining nature’s operation were no longer valid. The subatomic realm called for a new set of rules, or “mechanics,” which revealed the warp and woof of reality to be anything but a grand machine.

This was especially brought to light during discussions about a phenomenon known as the photoelectric effect. The photoelectric effect is a phenomenon whereby electrons are ejected from some materials when those materials interact with light. To explain this using the classic wave model of light required the assumption that the energy of the incoming wave (light) is absorbed by atoms of the material, that in turn transfer the energy to electrons. With sufficient energy, these electrons break their bonds and escape.

Such an explanation expects to see greater electron energies with greater light intensity.

For the classic wave model this meant that increasing the wave amplitude of the incoming light should result in higher electron energies. However, experiment revealed greater numbers of electrons, all with the same energy, were released when these materials were illuminated using light of greater amplitude. In other words, the energy of the escaping electrons remained independent of light intensity.

To explain the photoelectric effect Einstein considered the natural composition of light energy as comprising increments of energy termed “quanta”. These quanta are referred to as “photons”, or "light particles.” Photon energy is proportional to frequency, not amplitude. Therefore, regardless of light intensity, photons with a characteristic energy are delivered to the material. Electrons absorb the photon energy and escape with an energy equal to the bonding strength of that particular atom. If the intensity is increased, the number of escaping electrons increases. However, only by changing the

34 frequency of the incoming light (changing the color of light impacting the material surface) can a change be induced in the energy possessed by the escaping electrons.

Light of higher energy (higher frequency) will result in the release of more energetic electrons.

It should be noted here that the concept of energy quantization utilized by

Einstein in 1905 to explain the photoelectric effect was introduced by the German physicist Max Planck (1858-1947) around the turn of the twentieth century.47 Building on Planck’s work, Einstein came to conceptualize light in terms of particles, or perhaps more correctly, packets of energy. We recall that Newton, too, held that light consisted of particles or corpuscles. However, whereas Newton attempted to explain all phenomena involving light in terms of his corpuscular theory, Einstein recognized that some optical phenomena, such as diffraction, are diametrically opposed to the notion of light particles, while others, such as the photoelectric effect, cannot be explained in terms wave behavior. There exists, therefore, an inherent wave-particle duality or complementarity in nature, with respect to light phenomena, as well as the fundamental constituents of the atom. Over time experimental evidence has verified that the electron, discovered by J. J. Thomson (1856-1940), does not always behave as a particle, but rather it exhibits behavior that can only be explained by regarding the electron as a wave.

Electrons accelerated by an electromagnetic field produce electromagnetic . Photons are also emitted when nuclei, atoms, and molecules lose energy.

Exchanges between energy states of a nucleus, atom, or molecule result from the absorption, or from the emission, of photons with energies characteristic of the particular

47 Bohr, “The Structure of the Atom,” 12-13. 35

system. The frequency of radiation emitted in such an exchange varies according to the

emission source (atoms, nuclei, molecules), with the frequency (and thus the energy) of

the emitted radiation being equal to some multiple of the difference between energy

levels or states within the atom. In other words, energy does not exist in nature as a

continuum, but rather comes in incremental jumps – energy is quantized. According to

this view, a quantum of energy is emitted only when an electron drops from one energy

level to another. Conversely the absorption of a quantum of energy will cause an electron

to jump from one energy level to another. This contrasts with the classical notion of

particle-particle interaction, whereby one would expect the continual emission of energy

as the electron moves about in its atomic orbit – resulting in eventual collapse of atomic

stability (the electron would eventually run out of energy and fall to the nuclear surface).

The observed quantum behavior of physical reality rendered classical particle-

particle explanations useless. It is not possible, for example, to state with certainty the

position and velocity of a subatomic particle when the trajectory of that particle takes

incremental jumps or steps, rather than moving in one smooth continuous path.

Fundamentally this was the problem that arose with the advent of quantum physics. In

1927 Werner Heisenberg (1901-1976) articulated the inherent indeterminacy within the

fabric of reality itself. His enunciation, known as the Heisenberg ,

states that our knowledge of the world around us comes by way of photons as discrete

packets of information. The fact that these exist in finite increments prohibits us from

possessing the ability of infinite refinement of the data.48 In other words, we as observers

can receive no information more detailed, or with greater resolution, than that contained

48 Werner Heisenberg, The Physical Principles of the Quantum Theory (New York, New York: Dover Publications, 1949), 14. 36 in complete photons, a discontinuous change known as a “quantum jump”49 expressed by a value known as Planck’s Constant. Through the act of observing physical quantities at the subatomic level, the measuring device we must use naturally interacts with, and therefore influences, the object in question, thereby introducing a new element of uncertainty into the very systems we wish to objectively examine, something referred to as the “measurement problem.” Heisenberg characterized the measurement problem in the following way:

Since the device is necessarily described in terms of ; such a description contains all the uncertainties concerning the microscopic structure of the device which we know from thermodynamics, and since the device is connected with the rest of the world, it contains in fact the uncertainties of the microscopic structure of the whole world.50

In other words, the concept of “observation,” a crucial part of the scientific method, involves indeterminacy because one cannot determine “…other than arbitrarily, what objects are to be considered as part of the observed system and what as part of the observer’s apparatus.”51 This Bohr-Heisenberg view of a fundamental indeterminacy in physical reality came to known as the Copenhagen interpretation of quantum physics.

The Copenhagen interpretation became the prominent scientific view of quantum behavior. To gauge just how revolutionary this proposal seemed consider the reaction of

Einstein. He maintained “…his belief in the reality of an external world controlled by causal mechanisms that it was the subject of science to disclose.”52 Underlying this belief was the conviction that more complete knowledge of the laws governing subatomic behavior would result in a causal accounting of such behavior. It is ironic that Einstein, who was never able to accept quantum theory’s implication of this indeterminacy of

49 Ibid., 54-55. 50 Heisenberg, Physics and Philosophy: The Revolution in Modern Science, 53. 51 Heisenberg, The Physical Principles of the Quantum Theory, 64. 52 Brooke, Science and Religion, 449. 37 subatomic physical reality, contributed to quantum theory by way of his explanation of the photoelectric effect.

With the quantum view of nature and its lack of determinacy, the view of the world as machine dissipated. In its place a view developed that allows for potential and possibility. No longer is there a confidence in humanity’s ability to gather enough information to permit a complete, accurate description of reality, for reality itself cannot be known so intimately. Even if one could find someone to meet the criteria to be

Laplace’s “infinitely capable and industrious mathematician,” mentioned earlier, the mathematician still could not gather the data necessary to give a complete description of the world past, present, and future – for physical reality simply does not give up such data.

God and Quarks

The earliest modern physicists sought to pursue their scientific investigations out of a genuine curiosity to know more about how the world works, as is the case with modern physicists of today. Furthermore, they continued their work as people with varying beliefs and values. The work of science was not done in isolation, which meant that the earliest modern physicists were people whose beliefs, whether religious or non- religious, were integral to them as scientific investigators. At the same time what physics revealed about the world around them influenced these physicists’ beliefs. The work of scientific investigation continues today as a discipline that involves people with differing beliefs and values, both religious and non-religious. A survey of the personal writings of physicists of religious faith reveals their own thoughts and insights with regard to the

38 question of the relationship between science and faith, or more particularly between God and the world. Even some physicists who were not people of religious faith have had something to say with regard to the relation between physics and religion. As an example consider Niels Bohr, one of the founders of quantum physics. Some of Bohr’s biographers have claimed that he was an atheist, while other sources argued that it may be more correct to view him as an agnostic. His family upbringing included membership in the Lutheran . Heisenberg quoted Bohr as saying the following with regard to religious faith and science:

Still, religion is rather a different matter. I feel very much like Dirac: the idea of a personal God is foreign to me. But we ought to remember that religion uses language in quite a different way from science. The language of religion is more closely related to the language of poetry than to the language of science. True, we are inclined to think that science deals with information about objective facts, and poetry with subjective feelings. Hence we conclude that if religion does indeed deal with objective truths, it ought to adopt the same criteria of truth as science. But I myself find the division of the world into an objective and a subjective side much too arbitrary. The fact that through the ages have spoken in images, parables, and paradoxes means simply that there are no other ways of grasping the reality to which they refer. But that does not mean that it is not a genuine reality.53

This is not to say that science and theology search after answers to the same questions.

They are disciplines that address very different questions. This raises the question of whether they both address different aspects of something more fundamental. Bohr offered that the two disciplines may be complementary descriptions which, “…though they exclude one another, are needed to convey the rich possibilities flowing from man’s relationship with the central order.”54 Max Planck was a German physicist and a founder of quantum theory. In his words:

53 Werner Heisenberg, “Science and Religion,” Edge, June 3, 2007 https://www.edge.org/conversation/science-and-religion. 54 Ibid. 39

For these are wide spheres where they have absolutely nothing to do with each other. Thus all problems of ethics are outside of the field of natural science, whereas the dimensions of the universal constants are without relevance for religion.

On the other hand, religion and natural science do have a point of contact in the issue concerning the existence and nature of a supreme power ruling the world, and here the answers given by them are to a certain degree at least comparable. As we have seen, they are by no means mutually contradictory, but are in agreement, first of all, on the point that there exists a rational world order independent from man, and secondly, on the view that the character of this world order can never be directly known but can only be indirectly recognized or suspected. Religion employs in this connection its own characteristic symbols, while natural science uses founded on sense experiences.55

What is interesting here is not so much Bohr’s and Planck’s specific theological views,

but rather their consideration and recognition of the question of the relation between

science and theology. While Bohr indicated that science and theology exclude each

other, the sentiments of Planck, as well as many others before him, indicate that science

and theology can have some exchange or that dialog can occur between the two

disciplines. Planck maintained that “…both religion and natural science require a belief

in God for their activities, to the former he is the starting point, to the latter the goal of

every thought process.”56 The fact that science is a discipline limited to the study of the

physical universe using empirical evidence to provide explanations of how the world

works does not mean it rules out the claims of faith, nor does it mean that science can

directly verify faith claims. For science to be science, it must limit itself to those

phenomena whereby the scientific method can be applied. Modern science is not

designed to answer metaphysical questions. However, science can and does depend

critically “…upon values and presuppositions that are not the result of scientific

55 Max Planck, Scientific Autobiography and Other Papers, trans. Frank Gaynor (London: Williams & Norgate LTD. 1950), 182-183. 56 Ibid., 183. 40

discovery or testing.”57 Therefore they are not subject directly to evaluation by way of

empirical investigation.

As was the case with a number of the earliest modern physicists, some of the

physicists instrumental in contributing to subsequent revolutions in physics in the

twentieth-century were known to be people of faith. Michael Faraday (1791-1867), one

of the founders of , was a member of the Sandemanian of

Christianity. He gave attention to the relation between science and religious faith and,

like Planck, recognized that a clear distinction exists between the disciplines of science

and religion.58 (1831-1879), whose four equations are regarded as

some of the most influential equations in science, once wrote, “I think men of science as

well as other men need to learn from Christ, and I think Christians whose minds are

scientific are bound to study science that their view to the glory of God may be as

extensive as their being is capable of.”59 William Thompson (Lord , 1824-1907),

one of the founders of thermodynamics, was known as a devout Christian. After a series

of lectures on Christian apologetics that was given at University College, London in May

of 1903, The Times printed Kelvin’s response to the lecturer. Kelvin made some

corrections in his own hand to this newspaper article in which he wrote:

We only know God through his works, but we are absolutely forced by science to admit and to believe with absolute confidence in a Directive Power – in an influence other than physical, or dynamical, or electrical forces. Cicero, editor of Lucretius, denied that men and plants and animals could have come into existence by a fortuitous concourse of atoms…If you think strongly enough you will be forced by science to the belief in God,

57 Moritz, Science and Religion: Beyond Warfare and Toward Understanding, 67. 58 Colin Russell. “Science and Faith in the Life of Michael Faraday.” Faraday Paper no. 13, The Faraday Institute for Science and Religion, April 2007, http://faraday.stedmunds.cam.ac.uk/resources/Faraday%20Papers/Faraday%20Paper%2013%20Russell_E N.pdf. 59 Lewis Campbell, William Garnett, The Life of James Clerk Maxwell (London: Macmillan and Co., 1882), 404-405, https://archive.org/details/lifeofjamesclerk00camprich/page/404/mode/2up. 41

which is the foundation of all religion. You will find science not antagonistic, but helpful to religion.60

The key point here is that the question of the relationship of science to religious

faith was relevant during the birth of modern science, continued to be relevant throughout

the history of modern science, and remains relevant today. The consideration of the

relation between God and the world is a more specific question addressed within the

broader context of the relation between science and religious faith. What has changed

over time is not that the question of God’s relationship with the world is no longer valid

or important, or that modern physics has rendered the question irrelevant, but rather that

the question of God’s relationship to the world must now be considered within a new

context. This modern context came about due to the revolutions in twentieth-century

physics, and is now one in which science no longer views physical reality as a great

machine but rather something inherently possessing a measure of mystery and

indeterminacy. Furthermore, this indeterminacy in nature is not due to the limitations of

human knowledge, but rather to the very way that physical reality is structured. As an

example of how physics’ new view of physical reality affected the consideration of the

question of God and the world we can consider Arthur Holly Compton (1892-1962), the

physicist whose X-ray scattering experiments conclusively demonstrated the dual nature

of light as both a wave and a particle, and who was one of the key Manhattan Project

scientists. Compton was a Mennonite of whom Leona Woods, another Manhattan Project

physicist, once said “Arthur Compton and God were daily companions.”61 In an effort to

address the theological concept of human free will, Compton argued that Heisenberg’s

60 Silvanus P. Thompson, The Life of William Thompson Baron Kelvin of Largs (London: Macmillan and Company, 1910), 1098-1099, https://archive.org/details/lifeofwillthom02thomrich/page 1098-1099/n7. 61 Richard Rhodes, The Making of the Atomic Bomb (New York, London: Simon & Schuster, 1986), 363. 42

uncertainty principle extended in some way beyond the mere physical dimensions of the

atom into the human sphere and therefore confirmed the freedom of human will.62 While

respected for his scientific achievements, Compton was criticized by Niels Bohr as well

as others for what they regarded as his endeavor to manipulate the Heisenberg

Uncertainty Principle in this way.63 However, this effort on Compton’s part serves as one

example of how the long-standing question of God’s relationship to the world began to be

asked anew and in a completely new context with regard to the way natural science views

and understands physical reality. In recognition of this influence that modern science

exerted on theology along with its associated challenges, Nicholas Saunders stated:

The outstanding successes of the natural sciences over the past three hundred years have caused an intellectual shift that has challenged both biblical authority and the intellectual credibility of much received religious doctrine. Of all the challenges science has raised for theology, perhaps the most fundamental is that it has brought into question the doctrine of divine action.64

62 Ibid., 364. 63 Ibid. 64 Nicholas T. Saunders, “Does God Cheat at Dice? Divine Action and Quantum Possibilities,” Zygon 35, no. 3, (September 2000): 518. 43

CHAPTER 3

APPROACHING A VIEW OF GOD’S RELATIONSHIP TO THE WORLD

The new way modern physics modeled physical reality brought with it challenges,

especially challenges centering on answers to long-established questions. One way this

was manifested was that original metaphysical concerns of the early physicists came to

be re-expressed. As already mentioned, in light of these new challenges scientific

investigation into the nature of reality gave rise to religious and non-religious or anti-

religious approaches to these metaphysical concerns. All of these are possible options as

long as they grow out of consideration of the available evidence for “As with all

academic disciplines, that study must be open and honest, prepared to conform to the

discovered nature of reality and not shackled by prior prescription of what are the

acceptable outcomes of the enquiry.”65 Accordingly, the intent here will be to

acknowledge both religious and non-religious approaches, but to focus on the religious

by way of consideration of how the question of God’s relation to the world has come to

be addressed by contemporary physicists and theologians. Along with this are associated

concerns such as whether science and theology can dialogue with each other after what

some would argue were long periods of apparent estrangement, identifying the insights

that modern physics brings to the question of God and the world, and endeavoring to

characterize the nature of God’s relationship to the world. In light of the insights brought

about by modern physics, as opposed to classical physics, the question of the reality of

65 John Polkinghorne, Faith Science & Understanding (New Haven, : Press, 2000), 19. 44 divine influence and interaction within space-time regained significance. As previously noted, though it was widely criticized, Arthur Compton’s effort to extend Werner

Heisenberg’s uncertainty principle beyond the subatomic operations of nature and apply it to the realm of human behavior displayed what could very well be considered an awkward attempt to validate the biblical doctrine of free-will by appealing to a foundational principle of quantum physics.

The question of the relationship between God and the world goes to the heart of the relationship between the physical and the spiritual. The view of physical reality offered by mechanistic determinism did not view God as necessarily involved in any way with the fundamental dynamics of the universe or even as having any relevance at all. As a result, this way of thinking influenced many people of faith to invoke God in their own efforts to account for and explain phenomena or events that could not yet be fully understood and explained by science. This approach has come to be known as “God-of- the-gaps” theology. In the view of many theologians and scientists alike, “God-of-the- gaps” theology is inconsistent with the Christian understanding of the person and work of

God, as well as the insights of modern physics. While the modern-day scientist must acknowledge limitations in human understanding, and therefore the existence of epistemological gaps, such limitations and gaps cannot serve as a warrant for an argument for the presence and work of God. This is an argument made by high energy particle physicist and Anglican Priest John Polkinghorne:

Not that there are not many things which we do not understand….However it no longer seems plausible that there are scientific no-go areas, in which questions can be posed scientifically to which only a God of the Gaps could provide an answer. Scientific questions demand scientific answers and they seem to get them. As the theoretical chemist and devout Christian, Charles Coulson, briskly said, ‘when we come to the

45

scientifically unknown, our correct policy is not to rejoice because we have found God; it is to become better scientists.’66

Polkinghorne maintains that God-of-the-gaps theology is bad theology. Nevertheless, for many physics’ revolutionary view of the world leaves room for perceiving God at work within the ongoing life of the universe, not just at its birth or death (at the extreme boundaries of its oscillations). In other words, as physical reality came to be viewed in a completely new way, the question of God’s relation to the world also came to be considered in a new context. This resulted in restating some of the long-standing questions, as well as reformulating questions, or even asking different questions, about

God and the world.

The question of the relationship between the physical and the spiritual, and more particularly the question of gaining a better understanding of God and God’s relationship to the world, has been an integral part of the history of the birth of modern science.

Physicist and theologian (1923-2013) has characterized physics, the most fundamental of the sciences, as the first to be a systematic approach concerned with precision and using methods that other sciences chose to emulate.

Physics is the study of the basic structures and processes of change in matter and energy. Dealing with the lowest organizational levels, and using the most exact mathematical equations, it seems of all sciences furthest removed from the concerns of religion for life, mind, and human existence. But physics is of great historical and contemporary importance because it was the first science that was systematic and exact, and many of its assumptions were taken over by other sciences. Its methods were seen as ideals for other sciences to emulate. It exerted a major influence on philosophy and theology.67

More correctly, it might be argued that there has existed a mutual widespread influence of physics on philosophy and theology as well as an influence of philosophy and theology on physics as historians of science have pointed out.

66 John Polkinghorne, One World (Princeton, New Jersey: Press, 1986), 60. 67 Barbour, Religion and Science: Historical and Contemporary Issues, 165. 46

Today, it is commonly believed that faith – especially religious faith – only plays a negative role in science. The history of science however, shows this has generally not been the case. Historians of science have described numerous instances of scientific inspiration and motivation driven by key philosophical assumptions initially grounded in faith.68

More particularly, historians of science have argued that “The specific religious context

within which early modern science developed was the Christian faith as it was passed

down from the European Middle Ages.”69 In other words, the very process that

developed into the scientific method flourished in large measure due to the Judeo-

Christian view that the universe was, in Copernicus’ words “…wrought for us by a

supremely good and orderly Creator.”70 In this way it might be argued that the

significance of the question of God’s relationship with the world goes to the heart of how

modern science came to be and why modern science has been so successful in its mission

to more adequately describe the way the world around us works.

Today many scientists and theologians alike investigate the question of God’s

relationship to the world with the aim of arriving at a better understanding, or more

accurate knowledge, of the way things are. As practitioners they recognize that human

reason is corrigible. They also recognize an inherent relationship between science and

theology since both, in their own distinct ways, seek an understanding of certain truths

with respect to reality. This, then, facilitates relating Christian faith and natural science

through a consideration of whether and how God interacts with the world. Many

approaches have been taken with regard to addressing this question in light of the

revolutionary insights of modern physics concerning the nature of physical reality. In

68 Moritz, Science and Religion: Beyond Warfare and Toward Understanding, 55. 69 Ibid., 72. 70 Nancy R. Pearcy, Charles B. Thaxton, The Soul of Science: Christian Faith and Natural Philosophy (Wheaton, Illinois: Crossway Books, 1994), 25. 47

addition, the complexities of these insights, along with the variations in the explanations

of them, have contributed to greater variety in the approaches taken to the question of

God’s relationship to the world. Each approach has its major proponents who are

engaged with others in the discussion. Time and space will not permit a survey of each

of these approaches, nor is it my intent to try to do so here. However, in order to better

situate the question of God and the world within the broader historical context, it may

help to take a brief look at how varied these approaches are as well as the various ways in

which they have been characterized. Then we can move forward toward a more detailed

discussion of one particular approach that may hold promise.

Theological Models of God’s Relation to the World

Contemporary scholars have categorized these approaches in various ways, and

there seems to be no standard way to do so. However, as an early and premier scholar in

the field of science and religion, Ian G. Barbour, endeavored to situate the breadth and

depth of the discussion, along with all of the individual conversations, concerning God’s

relation to the world. Barbour gave two series of at the University of

Aberdeen from 1989-1991. The first of these was published as the book Religion in an

Age of Science (1990), which was subsequently revised and expanded as Religion and

Science: Historical and Contemporary Issues (1997). In 1999 Barbour was awarded the

Templeton Prize for Progress in Religion.71 He characterized the various approaches to

the question of God’s relationship to the world by ordering them into models.

71 “Ian G, Barbour, Winifred and Atherton Bean Professor Emeritus of Science, Technology and Society, ,” The Gifford Lectures: Over 100 Years of Lectures on Natural Theology, accessed July 3, 2018, https://www.giffordlectures.org/lecturers/ian-g-barbour. 48

He does this in part because models are integral to scientific investigation.

Barbour pointed out “…that in science there is no direct route by logical reasoning from data to theory. Theories arise in acts of creative imagination in which models often play a role. Here we are talking about conceptual or theoretical models, not scale models constructed in the laboratory, or logical or mathematical models, which are abstract and purely formal relationships.”72 Note here that he distinguishes between theoretical models and mathematical models. Barbour noted that one characteristic of conceptual scientific models is that they are analogical. As an example, the Bohr model of the atom uses one domain, planetary systems, to help explain the domain of atomic systems. In the analogy we find some general similarities that can help one better understand the atom, and yet many differences exist between these two realms. Even though much has been learned about the atomic and subatomic realms since Bohr offered his model, the model continues to be widely used today for explanations of the basics of atomic structure in introductory physics textbooks. Barbour also pointed out that conceptual scientific models help to extend theories. In other words “…it is often the model rather than the theory that suggests its application to new phenomena or new domains. It was the billiard ball model that suggested how the kinetic theory of gases might be applied to gas diffusion, viscosity, and heat conduction.”73 In this sense models at times seem to act as stand-alone entities. This brings us to the third characteristic that Barbour noted regarding conceptual scientific models, which is that they can be understood as units. As he understands it models cannot substitute for the fundamental language of science, which is mathematics. However, they may be more readily perceived and comprehended

72 Barbour, Religion and Science: Historical and Contemporary Issues, 116. 73 Ibid. 49 than more complex concepts or abstract mathematical equations. In this way models can be a good way to introduce the concepts and associated mathematics that constitutes the language of science.

Accordingly, Barbour categorized some of the key theological models relating to the question of God’s relation to the world. He held that “…models of the divine are crucial in the interpretation of .”74 Furthermore, he maintained that theoretical religious models have the same three fundamental characteristics as conceptual scientific models. With regard to the approaches to the question of God’s relationship to the world Barbour ordered these into eight models, acknowledging that each one has strengths as well as shortcomings. He referred to the first one as the

“monarchical model,” which primarily focuses on the transcendence, power, and sovereignty of God. In Barbour’s words, “This model was already present in the biblical view of God as Lord and King…Some parts of science are in keeping with this model: the awesome power of the , the contingency of the universe, the immense sweep of space and time, and the intricate order of nature.”75 A key difficulty that he identified with this model has to do with the effort to reconcile the apparent conflict between the model’s elaboration on and predestination with 1) evidence of human freedom, 2) the and suffering, 3) the elevation of masculine over feminine virtues, 4) religious , 5) an evolutionary view of the world, and 6) the presence of chance and necessity in nature.

He identified the second model as the “neo-Thomist model” of double agency.

This model focuses on primary and secondary causes operating in different realms.

74 Ibid., 119. 75 Ibid., 329. 50

Accordingly, this reinforces the role of natural causes while also recognizing these as indirectly predetermined in God’s plan. In this way God maintains and concurs with the natural order. For Barbour “…all the problems inherent in the concept of omnipotence are still present.”76 This means that specific divine initiatives (salvation by faith in

Christ, miracles, etc.) amount to being supernatural interventions into, or interruptions of, the natural order.

God as “determiner of indeterminacies” is the third model identified by Barbour.

He characterized this model as being consistent with the most common interpretation of quantum mechanics - the Copenhagen interpretation. According to this model God does not intervene in, or interrupt, the laws of nature, but rather causes one outcome among a range of potential outcomes at the quantum level of nature to occur. Though the outcome occurs at the quantum level (subatomic level) of nature, it must be amplified by some means in order to effect events at the macroscopic level, by way of such things as mutations, neural events, or chaotic systems. Therefore, one of the major concerns for this view is to provide an adequate mechanism, or mechanisms, whereby quantum occurrences are amplified in such a way as to have consequences in the macroscopic realm. In addition, if God controls all of nature’s indeterminacies, then this implies a divine determinism, which brings with it the same problems cited with the second model.

A chief problem with this third model for some is its assumption of a bottom-up in nature, which has typically been associated with reductionism,77 mentioned briefly in

Chapter One above.

76 Ibid., 330. 77 Ibid. 51

The fourth model is one about which Barbour stated “…seems to be a promising model.”78 He referred in this model to God as being the “communicator of information,” pointing to its reliance on the importance of information in science, and in particular information at the various levels of nature. With regard to this model he points out that

“The message carried in any is dependent on the wider context of interpretation and response.”79 It is important to note that Barbour seems to be using the term information to refer to the communication of a message, a prominent concept in physics, as well as other branches of science. Moreover, he points out that in this model the message communicated is dependent on a broad context that includes interpreting the message as well as responding to the message. This involves both the source and destination of the message, and is not merely a matter of information in the abstract. It is also of importance to Barbour to note that it has been argued that biblical support exists for this model in the concept of the divine Word (logos). Rather than a bottom-up causality identified as potentially problematic in the third model, this model points to

God’s action in the world as a top-down causality from higher to lower levels. More will be said about top-down causality later on.

It is worth taking time here to better introduce what is meant by reference to chaotic systems, mentioned in Barbour’s discussion of the third model above. The reason for doing this here is due to the unique terminology and the fact that it is going to be encountered in the forthcoming discussion of John Polkinghorne’s approach to developing a better understanding of God’s relation to the world. This fourth model in

Barbour’s list offers a way in which quantum events may be amplified to produce

78 Ibid. 79 Ibid. 52 macroscopic significance, which is important for Polkinghorne. He argues that quantum events may be amplified by way of being part of larger systems in nature that have come to be known as “chaotic systems.” Chaotic systems are macroscopic systems in nature that exhibit an extreme sensitivity to their initial conditions.80 Such sensitivity prevents chaotic systems from being able to be completely isolated from their ambient environments, meaning that they must be considered “…holistically, in their total overall context.”81 Another characteristic of such systems is that the initial conditions work to constrain the behaviors of these systems to a range of possible behaviors. When the behavior of chaotic systems is modeled mathematically, the range of possible solutions is referred to as a “strange attractor.”82 The term strange attractor denotes the characteristic of the equations’ solutions to tend toward a particular range of values. Accordingly, chaotic systems can exhibit a measure of order by way of this restricted range of possible outcomes. The result is that multiple potential future outcomes exist for such systems.

As applied to physical systems or processes, these outcomes may occur at the same energy level while differing only in their respective patterns, or the details with respect to their particular development.83 We will return to consider chaotic systems and strange attractors in more detail later on.

Barbour referred to the fifth model as the “kenotic model,” involving God’s purposeful self-limitation. For some this answers many of the objections to the monarchical model. God’s self-limitation makes allowance for human free will, the freedom of natural processes, and makes the question of evil and adversity more

80 Polkinghorne, Faith Science & Understanding, 120-121. 81 Ibid., 122. 82 Ibid., 121. 83 Ibid. 53 manageable. God’s act of self-limitation does not necessitate a limitation in God’s ultimate power, but rather aligns with the Christian notion of reconciliation and the suffering of Christ.84 In some ways Barbour saw the “kenotic model’ as moving closer toward the “process model,” which is the eighth model that he identified.

The sixth model in Barbour’s list is the model of “God as agent,” which is in keeping with the biblical notion of identifying God by way of actions and intentions. The problem with efforts to characterize this model has to do with the tendency of scholars to end up completely isolating religious and scientific language in their endeavors to articulate helpful distinctions between them. In fact, this tendency seems to be the predominant problem Barbour had with this model.85

Seventh on Barbour’s list is the model of “the world as God’s body,” with its strong emphasis on divine immanence. Barbour characterized this model by pointing out that “Advocates of this model say that the relation of God to the world is even closer than that of the human mind to the body, since God is aware of all that is and acts immediately and directly.”86 While this model has appeal for those who argue for ecological and social responsibility using theology, key problems with this model are that 1) it does not sufficiently allow for the freedom of God or of human agents in relation to each other, 2) nor does it adequately account for divine transcendence.

The eighth model cited by Barbour is the “process model,” which holds the view that God is a “creative participant” in creation. In Barbour’s words:

God is a creative participant in the cosmic community. God is like a teacher, leader, or parent. But God also provides the basic structures and novel possibilities for all other

84 Barbour, Religion and Science: Historical and Contemporary Issues, 330. 85 Ibid. 331. 86 Ibid. 54

members of the community. God alone is omniscient and everlasting, perfect in wisdom and love, and thus very different from all other participants.87

Process thought with regard to God and the world aligns with “…an ecological and evolutionary understanding of nature as a dynamic and open system, characterized by emergent levels of organization, activity, and experience.”88 Some participants in the discussion of God’s relationship to the world find this view to be in accord with the view of physical reality rendered by modern physics. Barbour himself argued that the process model is able to provide distinctive answers to the six problems that he identified in the monarchical model. Accordingly, he concluded that this model has fewer weaknesses than the others he described. However, it is necessary to be reminded that “…according to critical realism, all models are limited and partial, and none gives a complete or adequate picture of reality.”89

The question of God’s relation to the world has resulted in myriad discussions among those adhering to views far more complex and nuanced than the general characteristics cited for Barbour’s models here. Barbour endeavored to give us a way to categorize the views held with regard to God’s relation to the world because the views are too varied and nuanced to give a thorough rendering of them in a single work.

Barbour looked for commonalities among various families of views that he termed

“models.” In other words, within each of the models cited there exist nuanced views and diverse conceptualizations among those who generally adhere to that particular model.

Barbour’s efforts evidence development and progression regarding the question of God’s relation to the world in light of the new way that modern science has come to view the

87 Ibid. 88 Ibid. 89 Ibid. 55

world. The question of God’s relation to the world still remains but in new contexts and

new ways of understanding both God (theology) and the world (physics). In light of this

we can proceed to focus on one particular view of God’s relation to the world, and work

to better understand this view along with its nuances, perceived strengths and

weaknesses, and criticisms.

John Polkinghorne and the Communicator of Information Model

Likely the most prominent advocate for the “Communicator of Information

Model” of God’s relationship to the world is Dr. John Polkinghorne (b. 1930), a high-

energy particle physicist and Anglican priest who approaches the pursuit of knowledge

by beginning with science and proceeding to examine the realm of natural phenomena for

evidence of agreement and consistency with the claims of religious faith. This approach

of beginning with the particulars and then moving on to more general principles has led

Polkinghorne to characterize himself as a bottom-up thinker. In his words:

The world of thought divides into top-down thinkers, who place reliance upon general principles and pursue their clear and discriminating evaluation, and bottom-up thinkers, who feel it is safest to start in the basement of particularity and then generalize a little. Plato versus Aristotle one might say. As a physicist my sympathies are with the latter.90

The place to begin is with the particularities of the world, and Polkinghorne finds that the

world, with all of its individual components and processes, is infused with value, which

gives us a context and framework in which to conduct our investigations. In other words,

scientific investigation, wherever undertaken, is always value-laden.91

90 John Polkinghorne, Faith of a Physicist (Minneapolis Minnesota: Fortress Press, 1994), 11. 91 John Polknghorne, Belief in God in an Age of Science (New Haven, London: Yale University Press, 1998), 17. 56

Polkinghorne served for twenty-five years as Professor of

at Cambridge University, and was named as a Fellow of the Royal Society in 1974 for his

mathematical models used for calculating the paths of quantum (subatomic) particles. In

1979 he resigned his teaching position at Cambridge University and trained to become an

Anglican priest. Ordained in 1982 he served for five years as a parish priest in the

Church of . Subsequently in 1986 he was appointed fellow, dean, and chaplain

of Hall, Cambridge, and then later, in 1989, was appointed president of Queens’

College, Cambridge, from which he retired in 1996. Throughout his career in science

and theology, as well as ministry, Polkinghorne has written extensively on the question of

God’s relationship to the world. In 2002 he was awarded the “The for

Progress Toward Research or Discoveries about Spiritual Realities.”92 In his Gifford

Lectures given at the during 1993-1994, he confessed:

Throughout, my aim will be to seek an understanding based on a careful assessment of phenomena as the guide to reality. Just as I cannot regard science as merely an instrumentally successful manner of speaking which serves to get things done, so I cannot regard theology as merely concerned with a collection of stories which motivate an attitude to life. It must have its anchorage in the way things actually are, and the way they happen.93

Polkinghorne wove his Gifford Lectures around the phrases of the in such a

way as to render a better grasp of his personal understanding of God and physical reality,

as well as his view with regard to the question of God’s relationship to the world.

One stream of thought that we find in Polkinghorne has to do with the nature of

how we know and what we know. His approach involves the phrase “epistemology

models ontology,” found frequently throughout his writings. He is optimistic that

92 Templeton Prize, 2020. John C. Polkinghorne – 2002 Press Release. [online] Available at: https://www.templetonprize.org/laureate-sub/polkinghorne-press-release/ [Accessed 24 November 2020]. 93 Polkinghorne, Faith of a Physicist, 8. 57

accurate, or more accurate, knowledge of the world is achievable, at least within human

limitations. According to him this is evidenced in the natural sciences.

I believe that the advance of science is concerned not just with our ability to manipulate the physical world, but with our capacity to gain knowledge of its actual nature. In a word, I am a realist. Of course, such knowledge is to a degree partial and corrigible. Our attainment is verisimilitude, not absolute truth. Our method is the creative interpretation of experience, not rigorous deduction from it. Thus, I am a critical realist.94

Polkinghorne defines critical realism “…as being the attempt to maximize the correlation

between epistemological input and ontological belief.”95 There is hope in his view,

which comes in the form of a correlation between the realms of phenomena (appearance)

and noumena (reality). From the standpoint of his training and career as a professional

physicist, he maintains that the critical realism of natural science has its counterpart in

theology, which means that the search for knowledge and understanding does not apply

solely to the natural scientist; it is also true of the theologian. The routes taken by these

disciplines are different, and the questions asked of each are different (science typically

asks “how,” theology typically asks “why”), but the goal is the same – a more adequate

understanding of things as they are. Confidence for being able to do this has been

expressed by (1924-2006) Anglican theologian and biochemist when he

stated that “More pertinently to the present context, since the aim of a critical-realist

theology is to articulate intellectually and to formulate, by means of metaphor and model,

experiences of God, then it behooves such a theology to take seriously the critical-realist

perspective of the sciences on the natural, including the human, world.”96 Furthermore,

this desire to understand what is, and to understand it as it is, is in no way limited to the

94 Polkinghorne, Belief in God in an Age of Science, 104. 95 Ibid., 53. 96 Arthur Peacocke, Theology for a Scientific Age: Being and Becoming – Natural, Divine, and Human (Minneapolis, Minnesota: Fortress Press, 1993), 21. 58 trained professional. It is the desire of every believer to gain some understanding of that which he or she . The two are not totally separable, for “Theology has long known that one must believe in order to understand…and yet one must understand in order to believe.”97

The view that we are able to comprehend something of the world around us, rather than wrestle with ideas and concepts that may have no correlation to reality, is one that we find in the fabric of modern science. Albert Einstein wrote:

In speaking here concerning ‘comprehensibility’, the expression is used in its most modest sense. It implies: the production of some sort of order among sense impressions, this order being produced by the creation of general concepts, relations between these concepts, and by relations between the concepts and sense experience, these relations being determined in any possible manner. It is in this sense that the world of our sense experiences is comprehensible. The fact that it is comprehensible is a .98

Polkinghorne is confident that our knowledge of the world is not misleading or illusory.

Rather, it happens by way of a “maximum correlation” between the intellect and things in the world, which is why he offers hope that we can truly know something about the world around us. Furthermore, he uses scientific critical realism to articulate a view of theological critical realism, which is vital to what he wishes to do toward offering a viable framework for the modern thinker to use in considering the relation between God and the world. The result he aims for is correct thinking with regard to things as they are.

In other words, our knowledge can correspond to the reality we are trying to comprehend, but we need the confidence that comes from the correlation between the two. This is what he means by “epistemology models ontology.”

97 Polkinghorne, Belief in God in an Age of Science, 115. 98 Einstein, Out of My Later Years, 61. 59

With respect to the nature of physical reality, he identifies the “classic”

metaphysical options to be , idealism, and dualism, and seeks a metaphysical

proposal that overcomes the shortcomings of each. In particular he wishes to avoid

materialism’s devaluation of the mental, idealism’s devaluation of the physical, and

dualism’s inability to integrate the two.99 On first consideration this seems to be a

difficult challenge but Polkinghorne is optimistic in this regard. He is encouraged by the

potential of what he terms “dual-aspect monism,” which is a view of reality from the

perspective of complementarity. By this he means that reality, rather than primarily

existing as material or mental, actually exists with material and mental as extreme poles.

In his words: “There is only one stuff in the world (not two – the material and the

mental), but it can occur in two contrasting states (material and mental phases, a physicist

would say) which explain our perception of the difference between mind and matter.”100

The term “complementarity” is drawn from twentieth-century quantum physics, and is

useful here since “Complementarity, as the quantum physicists call this delicate behavior,

is the scientist’s equivalent of the theologian’s perichoresis, the mutual indwelling of

characteristics.”101 Polkinghorne uses an analogous way of articulating the mental-

material aspects of the world by reference to the wave-particle duality of the subatomic

realm. It would seem that the benefit for him, and those endeavoring to follow his

thought, is that quantum physics provides language that allows us to begin to

99 John Polkinghorne, “The Metaphysics of Divine Action” in Chaos and Complexity: Scientific Perspectives on Divine Action, eds. Robert John Russell, , and Arthur R. Peacocke (Vatican City State: Vatican Observatory, 1995; Berkeley, California: Center for Theology and the Natural Sciences, 1995), 154. 100 Polkinghorne, Faith of a Physicist, 21. 101 John Polkinghorne. Science and Creation (Boston, Massachusetts: Shambhala Publications, 1988), 70. 60

conceptualize something that is “…so counterintuitive in terms of common-sense

expectation that it cannot be reduced to a simple-minded objectivity.”102

This seems to provide Polkinghorne with an analogous way of speaking about the

nature of the reality physics seeks to know. In his words: “Needless to say, I cannot

solve the problem of how brain and mind relate to each other, but I look for a solution

along the lines of a dual-aspect monism, a complementary account of matter in

‘information’-bearing-pattern,….”103 That the coupling or connection between mind and

matter cannot be fully explicated should not keep us from looking to this

complementarity, this mind-matter duality (not dualism), as a way to begin to understand

the interaction of the spiritual and the material, and more particularly, of God’s

interaction with the world.

Another crucial stream of thought in Polkinghorne’s writing has to do with his

understanding of God, which naturally weighs heavily in the approach he takes toward

answering the question regarding God’s relation to the world. Like physicists before

him, Polkinghorne argues that the universe points to something beyond itself.

Furthermore, he offers that this something is God, who possesses aseity, meaning that

God is self-authenticating and therefore his nature needs no evaluation by any other

criteria and no explanation in terms of anything else. He maintains that God is not

merely one entity among others, but “Rather he is the source of all that is, the One

omnipresent to every human experience.”104 Therefore, God is not merely part of the

metaphysical dual-aspect monism just described, but rather his active will is the

102 Polkinghorne, “The Metaphysics of Divine Action,” 148. 103 Polkinghorne, Belief in God in an Age of Science, 49-50. 104 John Polkinghorne, Reason and Reality (Valley Forge, Pennsylvania: Trinity Press International, 1973), 56. 61

sustaining ground of created reality.105 What Polkinghorne wants to avoid though is the

risk of “…attenuating our understanding of the divine to a merely deistic notion of the

sustaining ground of all.”106 Instead, he offers that God’s relation to his creation means

that God responds to specific circumstances. He offers that “The nearest analogy in the

physical world would be a universal medium, such as the nineteenth-century aether, or

the twentieth-century quantum vacuum.”107 He maintains that God is personal, revealing

himself in ways “...that are appropriate to the individuality of circumstance.”108 For

Polkinghorne this means that God is not isolated from creation, but rather is its

“sustainer” (upholding the universe as the ground of all that is), “lawgiver” (imparting

freedom and giving purpose to the universe and to humanity), and “interactor” (involved

with what is continually going on in the universe and in the of individual people).109

According to Polkinghorne, something of the complexity of the divine nature is

suggested by Trinitarianism and dipolarity.

Each of the divine persons is to be conceived as possessing his eternal and temporal pole. It will be the temporal pole of the Son which is involved in the kenotic focusing of the infinite upon the finite, in the historic episode of the incarnation. The temporal poles of the Father and the Spirit would continue God’s rule over the general process of the world and his immanent working within it, without any suspension of divine providence in the early years of the first century.110

Polkinghorne appeals to a much more social conception of the divine Trinity than what

has traditionally been encouraged by the classical Christian of Western

theology. In defending this view, he relies on the work of Jürgen Moltmann (b. 1926)

105 Polkinghorne, Faith of a Physicist, 52-53. 106 Polkinghorne, Reason and Reality, 56. 107 Polkinghorne, Faith of a Physicist, 53. 108 Ibid., 54. 109 John Polkinghorne, “God’s Action in the World,” J.K. Russell Fellowship Lecture, 1990, 2-8, http://www.starcourse.org/jcp/action.html. 110 John Polkinghorne, Science and Providence: God’s Interaction with the World (Boston: Shambhala Publications, 1989), 100. 62

describing him as “…a vigorous defender of such a social approach.”111 Moltmann

focuses on the uniqueness of each person of the Trinity and perceives the Trinity by way

of the perichoresis of the divine persons. His understanding of the trinitarian God is by

way of the three divine persons in a mutual loving relationship. In his view the relation

between God and the world is a reciprocal relationship whereby God affects the world

and yet is affected by the world. God relates to the world in his own trinitarian

experience so that his changing experience of the world is a changing experience of

himself. It has been explained that “Moltmann abandons the traditional distinction

between the immanent and economic , between what God eternally is in himself

and how he acts outside himself in the world. The cross (and, by extension, the rest of

God’s history with the world) is internal to the divine trinitarian experience.”112 So it is

in the relationships with each other that the three are persons. The unity of God is

defined in terms of love, and as perichoresis it can open itself to and include the world

within itself. Moltmann stated that

God is unselfish love. Kenosis is the mystery of the Trinitarian God. By virtue of God’s unselfish love, God permeates all creatures and makes them alive. In this way God lives in the creation community and allows the community of all creatures to live in God. In reciprocal permeation everything that is exists and lives. The unselfish empathy of God awakens the sympathy of all creatures for each other. Perichoresis is also the mystery of the creation.113

We can glimpse Polkinghorne’s reliance on Moltmann in his emphasis on God’s

relationship to time. Polkinghorne argues that a timeless view of God’s relation to the

world runs the risk of denying the reality of becoming. God has a relation to time,

111 Ibid. 112 Richard Bauckham, “Jürgen Moltmann” in The Modern Theologians, ed. David F. Ford (Malden, Ma.: Blackwell Publishing, 1997), 217. 113 Jürgen Moltmann, “God is Unselfish Love” in The Emptying God: A Buddhist-Jewish-Christian Conversation, eds. John B. Cobb Jr and Christopher Ives, (Eugene, Oregon: Wipf & Stock Publishers, 1990), 121 63 making him immanent within it as well as transcendent beyond it. With respect to God and time Polkinghorne criticizes classical theism’s attempt to assure that God is uninfluenced by the temporal. He maintains that such a view of divine impassability can result in conceiving God in isolation from his creation.114 He finds support in the theology of Moltmann for God’s provision of a way for something other than God’s self to exist. As Moltmann’s expressed it:

The existence of a world outside of God is made possible by an inversion of God. This sets free a kind of ‘mystical primordial space’ into which God – issuing out of himself – can enter and in which he can manifest himself. ‘Where God withdraws himself from himself to himself, he can call something forth which is not divine essence or divine being’.115

In other words, in withdrawing into himself God brings about the emptiness, or what

Moltmann calls the “God-forsakenness,” in which God brings about his creation. This articulation brings to mind Barbour’s characterization of the “kenotic model”. Moltmann further points out that, “…time cannot be a category of eternity. It has to become a definition of created being in its difference from the eternal being of God.”116 In other words, time is part of creation and God’s relation to creation includes God’s relation to time. Moltmann maintains a trinitarian notion of creation whereby, “Creation exists in the Spirit, is moulded by the Son and is created by the Father. It is therefore from God, through God, and in God.”117 In his view this notion of creation binds together the concepts of divine transcendence and divine immanence. He warns that overemphasis on

God’s transcendence results in deism, while overemphasis of God’s immanence results in . He cited Newton, who he regarded as holding a view of God as deistic or

114 Polkinghorne, Faith of a Physicist, 61-62. 115 Jürgen Moltmann, God in Creation (London: SCM Press, 1985), 87. 116 Ibid., 113. 117 Ibid., 98. 64 very close to deistic, as an example of one who overemphasized divine transcendence leading to deistic belief, and Spinoza, who viewed “Nature or God” as one being, as an example of one who overemphasized divine immanence leading to pantheistic belief.

Moltmann seeks to avoid isolating God from the world or merely identifying God with the world. His view is one in which God is actively present in the world, is influenced by the world, and the world exists in God. This view has been termed . In his words, “In the panentheistic view, God, having created the world, also dwells in it, and conversely the world which he has created exists in him.”118 This notion is alluded to in

Polkinghorne’s acknowledgement wherein he said “I see panentheism as the eschatological of creation, not its present status.”119 He considers panentheism, the notion that the world is part of God but yet God exceeds the world, to be an inadequate view for achieving a proper balance between divine immanence and divine transcendence. For Polkinghorne, in panentheism’s view creation is the resulting multifaceted consequences that inevitably emanate from God’s fruitfulness or creativity.

However, the inherent defect of panentheism in his view is that it denies the human experience of a true otherness of the world from God.120 According to him, “There are distinctions between God and the world that Christian theology cannot afford to blur.

They lie at the root of the religious claim that meeting with God involves personal encounter, not just a communing with the cosmos.”121

While Polkinghorne understands panentheism to be different from , he finds some value in process theology thought. More specifically he points

118 Ibid. 119 Polkinghorne, Faith of a Physicist, 64. 120 Polkinghorne. Science and Creation, 53. 121 Polkinghorne, Science and Providence: God’s Interaction with the World, 16. 65 out that he believes “…there are very valuable insights contained within it, particularly in its emphasis on the dipolar nature of God and his consequent true engagement with time.”122 While others have often told him that he is a process theologian, Polkinghorne does not believe that process theology provides an adequate ground for hope, which he said is “…central to an understanding of what is involved in a Christian view of God’s reality.”123 He criticizes process theology’s tendency to focus on God as a “fellow- sufferer” who can empathize with us, while ignoring that God is the one in Revelation

21:3-4 who will be with us and wipe all tears away.124

Polkinghorne is an adherent of the view that God influences the world by way of

“top-down causality” whereby the whole influences the parts (downward emergence).

This may seem to be counterintuitive, as physics and the other hard sciences have tended to find their successes in bottom-up causality (upward emergence), which derives from active, lively interaction between the constituent parts of larger systems. However, he finds evidence for top-down causality in the human experience and therefore considers whether it may not be found elsewhere.

The experience of human agency seems totally different. It is the action of the whole person and so it would seem most appropriately to be described as top-down causality, the influence of the whole bringing about coherent activity of the parts. May not similar forms of top-down causality be found elsewhere, including God’s causal influence on the whole of creation?125

Polkinghorne evidences a measure of modesty when he acknowledges that the notion of top-down causality is not without problems, and he specifically identifies two. First is the recognition that top-down causality must be far ranging, much more so than

122 Polkinghorne, Faith of a Physicist, 65. 123 Ibid. 124 Ibid. 125 Polkinghorne, “The Metaphysics of Divine Action,” 151. 66

“…simply the generation of long-range order or the propagation of boundary effects.”126

In other words, since long-range order and correlation can be accounted for through the bottom-up approach of the scientific method, there must be something much more extensive or non-local with respect to top-down causality. The second problem is closely related to the first, and that is the notion that top-down causality must have room to operate. That is, there must be an intrinsic openness within the bottom-up descriptions of natural science. What he is aiming at here is openness in the structure of reality, or

“ontological gaps,” for top-down causality to work, as opposed to those gaps in human knowledge, or “epistemological gaps,” which science helps to fill.

Polkinghorne acknowledges that room for ontological gaps in the world may actually be found in those processes that occur at the level of the subatomic, the quantum processes, which he has referred to as “the basement of subatomic processes.”127

However, he does not propose that this is the mechanism by which divine agency occurs.

In his words “I do not suppose that either we or God interact with the world by the carefully calculated adjustment of the infinitesimal details of initial conditions so as to bring about a desired result.”128 He points out that the argument that God acts at the level of quantum indeterminacies (Barbour’s “determiner of indeterminacies” model) in order to bring about desired effects encounters some inherent difficulties. First is that the indeterminacies inherent in quantum events tend to cancel each other out when sufficiently large numbers of these events combine to describe behavior of physical systems at the humanly-perceptible level, the macroscopic level. Second is the fact that

126 Ibid. 127 Ibid., 152. 128 Ibid., 154. 67

quantum indeterminacies relate to unique events whereby a particular state is observed

and recorded. Such events are what is meant by the term “measurement” in quantum

theory. Such measurements tend to be episodic rather than continuous, and “…a God

who acted through being their determinator would also only be acting from time to time.

Such an episodic account of providential agency does not seem altogether satisfactory

theologically.”129 When one makes a measurement of a quantum event, one interacts

with and influences the event, and hence the result. This is part of the problem of

adequately explaining how quantum effects register clear, distinct results in the

macroscopic realm. Furthermore, it is not necessarily clear what is meant by the

terminology “quantum event.” As a high-energy particle physicist who routinely

researched quantum entities Polkinghorne pointed out that it is difficult to clearly know

what constitutes a quantum event.

Our perplexity stems in part from not knowing what meaning to attach to ‘quantum event’. The complex many-layered character of the flux of the physical world does not encourage the belief either that the phrase should refer simply to those occasional happenings that could be classified as measurements, or that it is to be reduced to something like the rather banal unfolding of the consequences of the Schrödinger equation.130

In other words, physical reality involves processes occurring on all scales where the ends

of the continuum lie at the microscopic and macroscopic endpoints. While physicists

often refer to realms of study, such as the quantum realm or the macroscopic realm, they

are in fact referring to a single world rather than two distinct worlds where the quantum is

distinguishable from the macroscopic. An adequate discussion of any particular real-

129 Ibid., 152-153. 130 John Polkinghorne, “Physical Process, Quantum Events, and Divine Agency” in Quantum Mechanics, eds. Robert John Russell, Philip Clayton, Kirk Wegter-McNelly, and John Polkinghorne (Vatican City State: Vatican Observatory, 2001; Berkeley, California: Center for Theology and the Natural Sciences, 2001), 187. 68 world event, then, would really require that the discourse be able to refer holistically to all levels of physical reality. However, we do not possess the ability to do this because of the measurement problem. If we endeavor to describe what transpires in a quantum event we have to realize that the term “transpires” can only refer to an observation that yields a particular result. Nothing can be said with regard to the state of affairs between two observations; such language really has no meaning in quantum theory. As soon as an observation is made, interaction occurs between the quantum object and the measuring device, and in turn the rest of the world.131 Polkinghorne’s statement above points out a paradox of quantum theory that Heisenberg expressed this way:

It has been stated in the beginning that the Copenhagen interpretation of quantum theory starts with a paradox. It starts from the fact that we describe our experiments in terms of classical physics and at the same time from the knowledge that these concepts do not fit nature accurately. The tension between these two starting points is the root of the statistical character of quantum theory.132

A third difficulty has to do with how quantum entities are able to effect a difference in the way events unfold in the macroscopic world, the realm in which people conduct their daily lives. As Polkinghorne characterizes it “…this can only be through an enhancement of their effect due to their being part of a much larger system which is extremely sensitive to the fine details of circumstance.”133 In other words, there must be some mechanism by which individual quantum events are amplified.134 Polkinghorne argues that the character of chaotic systems is such that quantum effects are enhanced, or amplified, to become part of larger systems that are sensitive to the initial conditions.135 Recall that during the discussion of Barbour’s “communicator of information” model he briefly mentioned

131 Heisenberg, Physics and Philosophy: The Revolution in Modern Science, 56. 132 Ibid., 44. 133 Polkinghorne, Faith Science & Understanding, 120-121. 134 Ibid., 120. 135 Ibid., 120. 69 chaotic systems. From the brief introduction offered earlier it may be recalled that a chaotic system does not develop in a purely haphazard manner, but rather exhibits a measure of order in that it is restricted to a range of possible behaviors, or patterns of behavior, known as a strange attractor.136 The term “strange attractor” comes from the mathematical modeling of these types of systems whereby the associated mathematical equations tend toward certain sets of solutions.

This means that there are real-world chaotic physical systems which will exhibit a range of potential patterns of behavior. Moreover, these behaviors are not possible to accurately predict unless the initial conditions are known with unlimited accuracy. In other words, chaotic systems exhibit a measure of order that lacks predictability in that they are limited to a range of patterns of behavior, yet it is not possible to determine which behavior will manifest without infinitely accurate knowledge of the initial conditions. Myriad patterns and behaviors in nature are cited as examples of chaotic systems and show up everywhere. Examples of chaotic systems in nature include the dynamics of mountain streams, cloud patterns, currents, effects of air turbulence, blood flow through blood vessels, branches of trees, leaf patterns on plants, orbits in the

Kuiper belt (the donut-shaped region of icy bodies beyond the orbit of Neptune), outbreaks of epidemics, population dynamics in ecosystems, and multiple other phenomena of the natural world. Chaotic behavior can also be seen in computations. For example, two computers will each round off irrational numbers (numbers that cannot be written as a ratio of two integers) with limited precision. Computing the same system

136 Ibid., 121. 70

with the same initial conditions on two different computers will eventually result in

completely different of the results.

As applied to real physical systems or processes, Polkinghorne argues that the

potential behaviors in a chaotic system may occur at the same energy level, while

differing only in their patterns.137 He has stated:

The openness that a chaotic system may be interpreted as possessing corresponds to the multiplicity of possibilities contained within this strange attractor, and any one of the motions that is actually executed can be understood as corresponding to an expression of the information specifying its detailed structure (‘this way, then that way, etc.’).138

Chaotic systems are extremely sensitive to their initial conditions, and have the

characteristic whereby “…two systems whose initial conditions lie arbitrarily close to

each other will diverge very rapidly, and after a limited period of time the behavior of the

two systems will show no correlation whatsoever anymore.”139 To give us some idea of

the level of the kind of sensitivity that Polkinghorne is talking about in real-world

macroscopic chaotic systems, which subsequently leads to freedom of outcomes,

Polkinghorne cites the following example:

Molecules in a gas behave, in many ways, like small colliding billiard balls. After only 10-10 , fifty or more collisions have taken place for each molecule. After even so few collisions the resulting outcome is so sensitive that it would be affected by the variation in the gravitational field due to an extra electron on the other side of the universe – the weakest force due to the smallest particle the furthest distance away!140

It is important that Polkinghorne argues that the resulting behaviors of real-world chaotic

systems may occur at the same energy level while differing only in their patterns of

behavior. This means that the amplification of quantum events by way of macroscopic

137 Ibid. 138 Ibid., 121-122. 139 Taede A. Smedes, “Is Our Universe Deterministic? Some Philosophical and Theological Reflections on an Elusive Topic,” Zygon 38, no. 4, (December 2003): 920. 140 Polkinghorne, Science and Providence: God’s Interaction with the World, 28-29. 71

chaotic behavior would not involve a violation of the law of conservation of matter and

energy, one of the most fundamental laws in physics. This is crucial for his view of

God’s interaction with the world. It is also important here to recognize that the

significance of the sensitivity manifested by chaotic systems distinguishes them as

needing to be treated holistically and points to their characteristic of being open to top-

down causality by way of the input of information regarding specific patterns of

behavior, otherwise known as the strange attractor.141

Therefore, the concept of information input is not necessarily tied to energetic

causality. According to Polkinghorne “If God acts in the world through influencing the

evolution of complex systems, he does not need to do so by the creative input of

energy.”142 He uses this to advocate that God’s interaction with the world may come

about by way of what he terms “active information” input. The resulting unpredictable

behavior means that there exists genuine openness in the macroscopic realm.

Furthermore, he argues that this order that lacks predictability is approximated by the

deterministic equations of physics conceived and articulated by physicists. Perhaps this

is partly what Eugene Wigner (1902-1995), the Hungarian-American theoretical physicist

who worked on the Manhattan Project, recognized when he wrote:

The first point is that the enormous usefulness of mathematics in the natural sciences is something bordering on the mysterious and that there is no rational explanation for it. Second, it is just this uncanny usefulness of mathematical concepts that raises the question of the uniqueness of our physical theories.143

141 Polkinghorne, “The Metaphysics of Divine Action,” 154. 142 Ibid., 32. 143 Eugene Wigner, “The Unreasonable Effectiveness of Mathematics in the Natural Sciences,” in Pure and Applied Mathematics 13, no. 1, (February 1960), https://www.dartmouth.edu/~matc/MathDrama/reading/Wigner.html. 72

That we are able “To express a law of physics in the form of a

means to collect a potentially infinite set of events into a single scheme, in the framework

of which every event, by being related to all other events, acquires a significance and is

explained.”144 This ability, referred to as “algorithmic compressibility”, is an aspect of

the “…uncanny usefulness of mathematical concepts”145 to which Wigner referred. This

also led him to say that “The miracle of the appropriateness of the language of

mathematics for the formulation of the laws of physics is a wonderful gift which we

neither understand nor deserve.”146 With regard to this wonderful gift Polkinghorne

maintains that mathematics is the natural language for formulating physical theory.

Therefore, he does not make a distinction between mathematical and theoretical models,

as did Barbour.147 In physics theoretical models are mathematical models. This is not to

say that Polkinghorne equates with pure mathematics. Rather it is an

assertion on his part “…that (with partial adequacy in a model and verisimilitude in a

theory) there is an isomorphism (corresponding identity of structure) between

mathematical patterns and physical patterns, which makes the former explanatory of the

latter.”148 In other words, mathematical equations are statements about relations between

physical quantities in the real world. This is another way of expressing the critical realist

approach of scientific investigation in the search for understanding how the world around

us works. What we learn about the world through the scientific endeavor comes about in

144 Michael Heller, “Chaos, Probability, and the Comprehensibility of the World” in Chaos and Complexity, Scientific Perspectives on Divine Action, eds. Robert John Russell, Nancey Murphy, and Arthur R. Peacocke (Vatican City State: Vatican Observatory, 1995; Berkeley, California: Center for Theology and the Natural Sciences, 1995), 109. 145 Eugene Wigner, “The Unreasonable Effectiveness of Mathematics in the Natural Sciences.” 146 Ibid., 9. 147 Barbour, 116. 148 Polkinghorne, Reason and Reality, 29. 73

part by way of perceptions gained through sense experiences. Critical realism “…holds

that this perception, though real, is indirect, and mediated through models or

analogies.”149

Emergence in general has traditionally been conceived as a process that moves in

one direction. It has been shown that physical reality is constituted of quantities that are

described by involving families of subatomic particles such as quarks,

leptons (i.e. electrons, muons, neutrinos), bosons (i.e. photons, gluons, gravitons),

hadrons (i.e. protons, neutrinos), etc. Furthermore, the sense has prevailed in modern

science that greater understanding of the fundamental constituents of physical reality will

lead to greater understanding of the complexities of physical reality wherein new and

different properties arise. Physics, then, readily describes bottom-up causality, whereby

these new properties come about by way of interaction of the constituent components of

the macroscopic systems. However, Polkinghorne makes the following argument:

Yet it is possible that if subatomic particles are not ‘more real’ than cells or persons, then they are not more fundamental either. It is possible that emergence is, in fact, a two-way process; that it would be conceptually valid and valuable to attempt to traverse the ladder of complexity in both directions, not only relating the higher to the lower but also the lower to the higher.150

This suggests a measure of a degree of reciprocity between levels of physical reality. He

argues that “The emergences which are truly interesting and truly puzzling are those

where there seems to be a qualitative change – the ‘coming to be’ of life or

."151 His motivation for exploring emergence in both directions comes in

149 Alister E. McGrath, Science & Religion: An Introduction (, UK: Blackwell Publishers, 1999), 66. 150 Ibid., 34-35. 151 John Polkinghorne, “The Laws of Nature and the Laws of Physics” in Quantum Cosmology and the Laws of Nature, Scientific Perspectives on Divine Action, eds. Robert John Russell, Nancey Murphy, and C. J. Isham (Vatican City State: Vatican Observatory, 1999; Berkeley, California: Center for Theology and the Natural Sciences, 1999), 430. 74

part from the study of complex dynamical systems modeled by mathematical “chaos

theory.” He points to the work of Jules Henri Poincaré (1854-1912), French theoretical

physicist, as leading to the development of chaos theory. Poincaré worked early in the

twentieth century during the same time in which quantum physics was being

developed.152 He worked to develop a partial solution to the “three-body problem” which

involves taking the initial positions and velocities of three point masses and then solving

for their subsequent motions using Newton’s laws of motion and universal gravitation.

Poincaré was specifically solving for the orbits of three planets. When he later

discovered that he had made a mistake and that his partial solution did not indicate stable

orbits for the planets, he also discovered that a very small change in his initial conditions

would lead to vastly different orbits. This discovery indirectly led to the development of

chaos theory. The lesson from mathematical chaos theory is that simple deterministic

equations can model behavior that becomes unpredictable as it evolves in time.

Explanation of this statement is probably best achieved by considering the following

example cited by Polkinghorne:

Suppose one compares calculations starting with x = 0.3 and with x = 0.3001, initial conditions which differ by less than one part in a thousand. For a few repetitions of the calculation they will keep roughly in step but quite soon the calculations will diverge from each other, giving totally different behaviors.153

It is worth noting here that physical systems are considered deterministic in that

they obey deterministic differential equations, equations that model the behavior of a

system as it evolves over time. The deterministic aspect of a system refers to the way

that it develops from to moment, where the present state of the system depends

152 John Polkinghorne, “The Quantum World” in Physics, Philosophy and Theology, eds. Robert John Russell, William R. Stoeger, S.J. and George V. Coyne (Vatican City State: Vatican Observatory, 2000; Berkeley, California: Center for Theology and the Natural Sciences, 2000), 334. 153 Polkinghorne, Reason and Reality, 36. 75

on the immediately past state in a well-determined way through physical laws. The

equations modeling a system are deterministic in that they describe its continuous

evolution over time. However, as mentioned above Poincaré found that it can happen

that very small differences in initial conditions will produce great differences in the final

phenomena as the system evolves through time. This means that it is not necessarily true

that knowing the state of a physical system and the associated physical law at one point in

time allows one to accurately predict the state of the same system at subsequent points in

time. As a result, the study of chaotic systems shows them to possess a unique

characteristic, which is that unless the observer is able to know the initial conditions of a

system with unlimited accuracy “…one can only project their behavior in a small way

into the future with any confidence. Beyond that they are intrinsically unpredictable.”154

More about this characteristic will be discussed later on as part of the examination of two

major critics of Polkinghorne.

As this applies to Polkinghorne’s view of God’s relation to the world it is difficult

to be very specific with regard to how this model can be said to functions in one’s daily

life. The reason for this is that divine action is hidden in the cloudiness of unpredictable

processes possessing so much sensitivity that various types of causality cannot be

completely distinguished and disentangled from one another.155 In Polkinghorne’s

words, “This intermingling of providential grace with the freedom of nature means that

divine action will not be demonstrable by experiment, though it may be discernible by the

intuition of faith.”156 This sounds slightly reminiscent of Newton’s dilemma mentioned

154 Ibid. 155 John Polkinghorne, Quarks, Chaos, & (New York, New York: Crossroad Publishing Company, 1994), 72. 156 Ibid. 76 earlier on. Therefore, while he does not cite specific examples Polkinghorne does give some indication regarding how his view applies to everyday life. He is careful to point out that some offer views of God’s relation to the world which include a view of divine control that is too tight, that God knows the future by controlling it. This could lead to the adoption of a view of God who is ultimately unaffected by his creatures. Instead

Polkinghorne argues that God improvises “…in response to his creatures’ free actions, but who is not ultimately thwarted by them.”157 According to him this means people may petition God for things such as healing during a time of illness or a change of weather in a time of drought. Polkinghorne disagrees with those who would argue that things like one’s health, the weather, etc. all merely follow their respective natural laws with little or no room for God. He responds by offering three considerations - one scientific, one human, and one religious. The first consideration is a reminder that modern physics does not describe the world this way any longer. The view that the world is a machine on a grand scale running with clocklike precision regards God as the unseen mover, uninvolved in the day-to-day lives of individuals. Instead, modern physics describes a world more subtle and open than this, a world that is free and capable of becoming.

Polkinghorne refers to this as “free process”, which is analogous to human free will. The second consideration is a reminder that while modern science has dispensed with mechanistic determinism as the predominant view of the world, people have always known that they are not merely automatons. In other words, it did not require the advances in scientific reasoning for people to know they have the ability to will, to choose this way or that. As a part of the physical world people experience the ability to

157 Polkinghorne, Science and Providence: God’s Interaction with the World, 98. 77 maneuver within the processes of the world rather than to be wholly controlled by them.

Polkinghorne asks, “If we can act in a world that is sufficiently open to the future for this to be possible, may it not also be the case that God can act in it also?”158 Free will is a fundamental biblical doctrine, as well as a concept that is included in various forms in major religions in the world. The final consideration looks to major religions that use personal terminology to refer to God. Even though human reasoning and language fall short of the task of describing God, Polkinghorne makes the case that it is “…personal language that seems the least inadequate for this purpose. God is thought of as ‘Father’ not as ‘Force’.”159 By this he means to emphasize that God cares for the individual, that

God is one who takes a personal interest in his creatures. Therefore, as a personal interaction between a person and God prayer is efficacious. In Polkinghorne’s words:

When we pray…we offer our will to be aligned to the divine will. I believe that when this alignment takes place, things become possible that are not possible when human and divine wills are at cross purposes. Therefore, prayer is genuinely instrumental. It genuinely changes the world.160

It is important to keep in mind that there exists a certain amount of mystery and perplexity with regard to correct thinking about the nature of the relationship between quantum systems and macroscopic chaotic systems. That this is not a settled issue is recognized and acknowledged by Polkinghorne. One of the reasons he relies on chaotic behavior is that quantum effects need to be amplified in order to bring about large scale effects in the macroscopic realm. This is directly due to the measurement problem, which concerns the effort to understand how events in the quantum realm register distinct results when a macroscopic-sized measuring apparatus is used. Measurement events can

158 Polkinghorne, Quarks, Chaos, & Christianity, 64. 159 Ibid. 160 Ibid., 75. 78

involve processes, such as radioactive decay or gene mutation, where nuclear or atomic

events produce macroscopic consequences. However, generally measurement events are

recognized as episodic in nature and therefore, per Polkinghorne, would not seem to hold

much potential for describing or modeling the openness and freedom characteristic of

agency.161 Because agency involves freedom he appeals to a “…much more generalized

notion of ‘the amplification of quantum uncertainties’.”162 A second reason for

Polkinghorne’s reliance on chaotic systems has to do with their holistic character deriving

from extreme sensitivity to the smallest triggers. Rather than implying that such systems

must be considered at the local level, that is, the level of their initial conditions where the

smallest fluctuations occur, such sensitivity means that such systems must be treated

holistically, as they cannot be isolated from the impact of their total environment.163

While he acknowledges the deterministic basis of the mathematics used for

describing chaotic behavior in physical systems and processes, Polkinghorne believes

real-world chaos to be indeterminate. He proposes that the deterministic simplicity of

mathematical chaos theory is useful for modeling a world that is inherently open and

unpredictable by nature.164 As emphasized earlier, he relies on the recent recognition in

science “…that many dynamical systems – physical, chemical, biological, and indeed

neurological – that are governed by non-linear dynamical equations can become

161 John Polkinghorne, “Physical Process, Quantum Events, and Divine Agency” in Quantum Mechanics, eds. Robert John Russell, Philip Clayton, Kirk Wegter-McNelly, and John Polkinghorne (Vatican City State: Vatican Observatory, 2001; Berkeley, California: Center for Theology and the Natural Sciences, 2001), 189. 162 Ibid. 163 Polkinghorne, Belief in God in an Age of Science, 62. 164 John Polkinghorne, Beyond Science: The Wider Human Context (Great Britain: Cambridge University Press, 1996), 71-72. 79

unpredictable in their macroscopically observable behavior.”165 With such systems

mathematicians found that modern computers can be used to build up solutions in

piecemeal fashion, resulting in the manifestation of unique behaviors. By varying a key

controlling parameter a single unique solution first presents which seems normal and

well-behaved from a deterministic standpoint. Subsequently, however, a critical value is

reached whereby solutions bifurcate into two possibilities, either of which may first occur

after the critical point is reached. However, which solution will result is not predictable.

Subsequently the system can flip or oscillate between the possibilities, bifurcate into

more possible solutions, or begin to exhibit erratic behavior. It has been explained that

such chaotic behavior in systems is “…truly unpredictable – there never will be

sufficiently accurate knowledge of the parameters prevailing during the fluctuations

(especially if it proves necessary to take into account quantum effects) to predict

absolutely which way the system will go.”166 If physical systems in the world

demonstrate such openness, this is crucial for the question of God’s relation to the world,

in particular for the question “How does God interact with and effect change in the

world?” This has implications for a view concerning God’s activity in the world that is

acceptable from a scientific perspective, as fluctuations in chaotic systems provide a

mechanism for small effects (such as the results of quantum events) to manifest

themselves at the macroscopic level. Recall that Polkinghorne is not arguing that God

acts to influence quantum events, as is the case in Barbour’s “determiner of

indeterminacies” model above. Rather he is arguing that analogous to human agency

165 Arthur Peacocke, Theology for a Scientific Age: Being and Becoming – Natural, Divine, and Human (Minneapolis: Fortress Press, 1993), 50. 166 Ibid., 51. 80 where the whole influences the parts, God acts to influence events in the natural world through the openness inherent in its many systems and processes. With regard to his approach he argues that it is a misrepresentation of his “…ideas to suggest that they imply that agency arises from the local manipulation of either boundary conditions or microscopic processes, either by or God.”167 If God inputs information (active information input) for directing such systems, then such input of information might provide for a better, more adequate conceptualization of the way in which God’s active involvement in the world takes place, otherwise known as the “causal joint” by those involved in the discussion of God and the world.

167 Polkinghorne, Faith Science & Understanding, 101. 81

CHAPTER 4

CRITICISMS OF POLKINGHORNE’S “COMMUNICATOR OF INFORMATION”

MODEL

While Polkinghorne, like others wrestling with the question of God’s relationship to the world, has his detractors who argue for other approaches, time and space will require the focus here to remain on specific criticisms of his use of chaos theory as a mechanism for amplifying quantum effects and providing for the openness necessary for agency at the level of nature in which people live their daily lives, the macroscopic level.

Polkinghorne has been criticized on a variety of issues, but it seems that those dealing directly with his use of chaos theory get to the very heart of his view of God’s relation to the world. They also steer the discussion to the heart of the great mystery of the link between the quantum and macroscopic levels of physical reality. Two key participants in the ongoing discussion of God’s relation to the world focus particular attention on

Polkinghorne’s use of chaos theory. While both argue that physical reality at the macroscopic level does not possess and display the openness that Polkinghorne claims it does, they vary somewhat in the way each focuses on different aspects of how chaos theory specifically works in Polkinghorne’s view. One of these, , devotes more attention to criticizing Polkinghorne’s claim that in his view of God and the world, divine interaction with the world does not break the law of energy conservation.

The second critic considered here, Nicholas T. Saunders, is more directly concerned with

Polkinghorne’s motivation for using chaos theory, and therefore whether he is really able to use the theory to model physical reality.

82

Wildman’s Criticism of Polkinghorne

One form of this criticism argues that physical systems of the real world cannot

really be indeterminate because these systems are described and modeled by mathematics

grounded in deterministic laws and principles. A major critic who holds to this view is

Wesley Wildman (b. 1961), theologian, philosopher, and ethicist who teaches at Boston

University. Over the years Wildman has been a participant in discussions as part of the

Divine Action Project (DAP). This program was a joint effort sponsored by the Vatican

Observatory and the Center for Theology and the Natural Sciences (CTNS) initiated in

1988 with the aim to encourage discussion and support research on the question of God’s

action in the world, as one theological issue which most directly presupposes facts about

the physical world and its governing laws.168 While the question of God’s relation to the

world is ongoing, the DAP as a formal project ran through 2003. Along with neurologist

Patrick McNamara, Wildman also cofounded the Institute for Biocultural Study of

Religion (IBSR), a non-profit organization dedicated to investigating the links between

biology, culture, and religion. In addition, he is the founding director of the Liberal

Evangelical Project, an organization that claims to identify with Christians possessing

both liberal and evangelical instincts, and that works to provide resources and

encouragement to this group of what it terms “radical moderates.”

Wildman maintains a view of God and physical reality very different than that of

Polkinghorne.

My view of ultimate reality is quite different from Polkinghorne’s but I admire the apologetic and evangelical impulses in his theological work and support his efforts to

168 Wesley J. Wildman, “The Divine Action Project, 1988-2003,” Theology and Science 2, no. 1, (2004): 33. 83

push human truth-seeking ingenuity as far as possible in this direction. A mystical theologian of my sort is happy to see such projects advance, both for the light they may shed on the God-world relationship and for the testimony that their ultimate failure makes, namely, that ultimate reality, which is infinitely intimate, also infinitely surpasses human reason.169

Wildman personally advocates for a theological view of God as ground-of-being, as

evidenced by the following statement: “I urge battle-weary determinate-entity theists to

look over at the intellectuals gathered around the ancient idea of God as ground of being,

the power and creativity of the structured flows of nature, the ontological spring of matter

and value.”170 Ground-of-being theism is actually a family of theological perspectives

which have two negations in common, “…they deny that ultimate reality is a determinate

entity, and they deny that the universe is ontologically self-explanatory.”171 Therefore,

Wildman makes a distinction between the family of ground-of-being theology and those

which identify God as a determinate being of some kind. Wildman’s own

characterization of ground-of-being theism is as follows:

Ground-of-being theism only confirms the process theologian’s suspicion that GodC is religiously useless. At least the determinate-entity theists make an effort to preserve religious relevance by insisting that GodC is good in a humanly recognizable way! Ground-of-being theists basically accept the process theist’s analysis of ultimate reality as conceptually incomprehensible and morally impenetrable but then call it “God” anyway.172

169 Wesley J. Wildman, “Further Reflections on ‘The Divine Action Project’,” Theology and Science 3, no.1, (2005): 72. 170 Wesley J. Wildman, “Incongruous Goodness, Perilous Beauty, Disconcerting Truth: Ultimate reality and Suffering in Nature,” accessed June 2, 2020, 293 http://people.bu.edu/wwildman/media/docs/Wildman_2007_Incongruous_Goodness_prepub.pdf. 171 Wesley J. Wildman, “Ground-of-Being-Theologies” In The Oxford Handbook of Religion and Science, ed. Philip Clayton, September 2009, https://www.oxfordhandbooks.com/view/10.1093/oxfordhb/9780199543656.001.0001/oxfordhb- 9780199543656-e-37. 172 Wesley J. Wildman, “Incongruous Goodness, Perilous Beauty, Disconcerting Truth: Ultimate Reality and Suffering in Nature,” 276. Note that the use of the subscript “c” with “God” is a purposeful effort to refer to divine creativity. 84

Dave Rohr is a Ph.D. candidate in the religion and science track of the Graduate Religion

Program at Boston University. He offers a more nuanced characterization of Wildman’s

theological perspective. More specifically he argues that “Wildman's theological

perspective can be characterized as a mystical ground of being theism, an axiologically-

oriented religious , and an apophatic celebration of creative freedom before

ultimate emptiness.”173

Accepting this more nuanced view from someone who has studied and worked

with Wildman, it may be useful here to briefly characterize . This is

a view which melds the naturalist worldview with sentiments, perceptions, and values

typically associated with religion. Two central aspects of religious naturalism are 1)

adherence to naturalism as a worldview and 2) appreciation for the religious. Naturalism

is a view of how and why things happen in the world and holds that the universe is a

closed system that is not open to outside influence, either by a transcendent God (for

none exists) or by autonomous human beings (for they are a part of the larger whole

uniformity of the universe).174 The fact that religious naturalism involves an appreciation

for the religious does not mean the term “religious” is used in the traditional sense.

Instead, religious naturalists use it in a generic manner, whereby the religious or spiritual

sense is associated with the feelings of awe and mystery one experiences in response to

the grandeur of the scope and power of the universe. For the religious naturalist, then,

nature becomes the core focus and narrative of religious attention. Accordingly, religious

aspects of the world, those things that promote a set of beliefs and attitudes, are

173 Dave Rohr, “Beyond Dead Silence: The Need for Cataphatic Affirmation Within Wesley Wildman's Apophatic Religious Naturalism,” Religious Naturalism Resources, accessed on June 2, 2020 http://people.bu.edu/wwildman/relnat/. 174 James W. Sire, The Universe Next Door (Downers Grove, Illinois: Intervarsity Press, 2009), 70. 85

appreciated within a naturalistic framework. Religion is therefore a phenomenon

resulting from social, cultural, and evolutionary factors.

It has been argued that the roots of religious naturalism go at least as far back as

Baruch Spinoza (1632-1677).175 In his Ethics Spinoza argued that “All things are made

through the power of God. Because the power of nature is nothing but the power of God

itself, it is certain that insofar as we are ignorant of natural causes, we do not understand

God’s power.”176 In his view “Nature or God is one being, of which infinite attributes are

said, and which contains in itself all essences of created things.”177 In twentieth-century

physics Spinoza’s name shows up in the life of Albert Einstein. Spinoza was prominent

in the personal belief of Einstein. After his arrival to the United States, and perhaps in

part due to the lack of understanding and arguments over his theories of relativity (special

and general), Einstein was labeled by some as an atheist. In part due to these concerns,

Rabbi Herbert Goldstein of New York sent a cable to Einstein asking “’Do you believe in

God?’ Einstein cabled back: ‘I believe in Spinoza’s God, who reveals himself in the

harmony of all being, not in a God who concerns himself with the fate and actions of

men’.”178 Einstein actually considered the idea of a personal God to be problematic and

even stated that “The main source of the present-day conflicts between the spheres of

religion and of science lies in this concept of a personal God.”179 Einstein’s answer to

Rabbi Goldstein may have alleviated some of the concern about his being an atheist

175 Jerome A. Stone, “Religious Naturalism and the Religion-Science Dialog: A Minimalist View,” Zygon 37, no. 2, (June 2002): 381. 176 Benedict de Spinoza, A Spinoza Reader: The Ethics and Other Works, ed. and trans. Edwin Curley (Princeton, New Jersey: Princeton University Press, 1994), 15-16. 177 Ibid., 59. 178 Paul Arthur Schilpp ed., Albert Einstein, Philosopher-Scientist (La Salle, Illinois: Open Court Publishing Co. 1970), 103. 179 Albert Einstein, Out of My Later Years (New York, New York: Citadel Press, 1984), 27. 86 because over the years Einstein has often been cited as the example of a believing scientist, and even a devout Jew.

In Wildman’s view it is problematic to attempt to connect the appearance of purpose and design in nature with the notion of God’s action in the world by way of arguing that the appearance of purpose and design in the universe indicates a genuine teleology that is part of God’s plan. He argues that there exists a metaphysical ambiguity that complicates matters too much to be able to establish such clear-cut connections between order and structure in the universe and God’s relation to the world as one who intentionally interacts with it. As mentioned above, Wildman prefers to speak of God as the ground of being. Therefore, he argues that belief in a God who intentionally acts in the daily lives of people complicates the question of God’s relation to the world, with particular reference to the question. For him this is so extreme that there is justification, on moral and theological grounds, for rejecting the notion of

God’s intentional interaction with the world.180 This places Wildman in the position of discussing God’s relation to the world while at the same time holding to a view whereby he does not advocate for belief in intentional divine action in the world.181 Therefore, the question of God’s relation to the world is set in a completely different context for

Wildman than for Polkinghorne. While there is great divergence with regard to how each one views God, there is perhaps more room for agreement between them when it comes to how each views physical reality and how it is modeled and approximated by the language of mathematics.

180 Wesley J. Wildman, “The Divine Action Project, 1988-2003,” 32. 181 Ibid., 35. 87

With regard to the behavior of physical systems Wildman questions whether it is valid to apply mathematical descriptions of chaotic behavior to real-world physical systems and processes.182 He argues that the epistemic limitations and deterministic framework of chaos theory must be taken seriously. Furthermore, he believes these are inconsistent with each other. In his words, “…chaos in the strict sense is a mathematical abstraction not directly relevant to the physical world.”183 In other words, the deterministic basis of the mathematics of chaos theory does not actually approximate indeterminate real- world physical systems. Still, he acknowledges that quantum mechanics yields enough open questions that there remains hope for the discovery of a new kind of chaos theory.184

This does not necessarily imply that Wildman is referring directly to a new theory of chaos in the quantum realm (referred to as “quantum chaology”) because he regards quantum chaology to be what he calls a “vain hope.” He has characterized his criticism of

Polkinghorne through the following statement:

Human freedom and moral responsibility, as well as some interpretations of quantum measurement, strongly suggest ontological indeterminism, and chaos theory does not block these familiar considerations. Inspiration for defense of ontological indeterminism arises far more naturally from these springs than from the deterministic well of chaos theory. However, because the inspiration is there, the central hypothesis of Polkinghorne’s view, that the laws of nature are approximations, remains perfectly feasible. Though the concept of active information in chaotic attractors seems too dependent on determinism and on the reality of chaos in nature to work in its current form, I have argued that these considerations serve merely to constrain the ongoing development of Polkinghorne’s view.185

There are two main criticisms that Wildman levels against Polkinghorne. The first of these has to do with Polkinghorne’s notion that God’s interaction with the physical universe does not add energy to a physical system. In other words, Wildman

182 Ibid., 47. 183 Ibid., 49. 184 Ibid., 48-49. 185 Ibid., 49-50. 88

disputes Polkinghorne’s claim that his articulation of the “communicator of information”

model allows that God’s interaction with the world does not violate the law of

. According to Wildman, this claim “…depends on the infinite

closeness of orbits in chaotic attractors; a characteristic of mathematical chaos that

Polkinghorne hopes exists also in the physical world.”186 The problem with the “infinite

closeness of orbits” to which Wildman refers has to do with “infinite intricacy,” one

characteristic of strange attractors, which is a product of the mathematical chaos. Infinite

intricacy refers to the fact that mathematically the strange attractor is created by repeated

iterations, without end, of the same pattern. This means that if one were to graphically or

visually display these repeated iterations an image of the strange attractor would appear,

which is referred to as a fractal. However much one zooms onto a fractal image one will

always encounter the same pattern with which one started. This is another characteristic

of strange attractors that is referred to as “zoom symmetry.” While infinite iterations

may be possible in mathematics they are not regarded as characteristic of real physical

systems. The pattern formed, a strange attractor, yields the restricted range of constraints

for the numerical values in the mathematical field of dynamic systems. This means there

is a restricted range of outcomes or behaviors. Strange attractors exist in nature, and

because they appear similar at all levels of magnification, they are often regarded as

infinitely complex. Wildman argues that real physical systems do not really possess

genuine openness because it is not possible to have infinite iterations of a physical pattern

in nature. Therefore, he does not believe there exists what he calls “…a perfect match

186 Wesley J. Wildman, “A Counter Response on ‘The Divine Action Project’,” Theology and Science 3, no.1, (2005): 18. 89 between the mathematics of chaos theory and the dynamics of complex physical systems.”187

Furthermore, he maintains that if such a match did exist the notion of action without the addition of energy into physical systems (zero-energy action) does not make sense. What may need some clarification with regard to this point is whether Wildman means that it does not make sense to have active information input without the creation of added energy, or whether he means that action along one path of the strange attractor rather than another, without there being a change in energy, does not make sense.

However, he has stated:

Perhaps Polkinghorne should surrender the idea of zero-energy divine action. This would make God so much like another creature in mode of action that we might judge the theological cost to be too high, even for personalist theists such as Polkinghorne, whose tolerance for relatively literal language about God as personal and intentional is higher than for some other theologians.188

The second criticism of Polkinghorne by Wildman has to do whether there exists an “…’extended correspondence principle’ that can give confidence that mathematical chaos tells us something about nature.”189 There are correspondence principles in physics that inform us about the behavior of various physical systems. For example, there exists a correspondence principle informing us that the predictions of Einstein’s theory of special relativity tend toward the predictions of Newton when the relative involved are very small compared to the speed of light. Another, more interesting example is the correspondence that exists between the quantum and classical realms of physical reality. For both of these realms there exist verifiable theories and their

187 Wildman, “Further Reflections on ‘The Divine Action Project’,” 72. 188 Ibid. 189 Wildman, “A Counter Response on ‘The Divine Action Project’,” 19. 90 predictions can be and have been tested experimentally. Still, the connection between these two realms remains obscure, with behaviors in the quantum realm that seem to be counterintuitive to human experience. This may very well contribute to greater emphasis on the question of how the world in which we live our daily lives emerges from the behavior of entities in the quantum realm. This lies at the heart of the measurement problem, and also becomes the foundational question with regard to the amplification of quantum results by way of physical chaos. Wildman is skeptical that any correspondence exists between mathematical chaos theory and the behavior of physical systems in nature.

He is not necessarily entirely pessimistic about contemporary theology’s endeavors to address the question of God’s relationship to the world. In fact, he believes that with regard to this question current theological theories “…are as strong as they have been at any time since Hume and Kant, and this is largely due to the contributions of the

DAP.”190

Saunders’ Criticism of Polkinghorne

Another form of the criticism of Polkinghorne’s reliance on chaos theory as an answer to amplifying quantum events as well as providing indeterminacy at the macroscopic level has to do with the motivation behind this move on his part. A key critic of

Polkinghorne who takes this approach is Nicholas Saunders, theologian, physicist, and barrister. David Hoyle, Dean and Head of the Bristol Cathedral Foundation was a key influence for Saunders’ initial interest in theology. Furthermore, his academic advisor,

Fraser Watts, was influential in helping Saunders develop a greater interest in the

190 Wildman, “The Divine Action Project, 1988-2003,” 63. 91

relationship between theology and science. Watts is an ordained minister of the Church of

England and an academician who has worked in the areas of clinical psychology, natural

science, and theology. As a theologian and physicist Saunders has worked with the Theory

of Condensed Matter Group at the Cavendish Laboratory in Cambridge, England as well as

scientific colleagues at the European Laboratory for Particle Physics (CERN) in Geneva,

Switzerland.191 There he worked on the design of some of the instruments for the Large

Hadron Collider. Therefore, he approaches the question of God’s relationship to the world

from a context that is similar in many ways to Polkinghorne’s, as a physicist who is a

bottom up thinker, and yet as a theologian from within the tradition of the Church of

England, who holds that God is intentionally involved and active in the world.

As a scientist and theologian Saunders maintains that methodological approaches are

integral to any meaningful discussion regarding issues relating to the interface between

science and theology, and begins his own framework for addressing the question of God’s

relationship with the world with the assumption that theoretically science and theology are

reconcilable. Saunders approaches this discussion with “…the presumption that God exists

and is active in the natural world in a continuing and particular sense (i.e. that God performs

special divine actions in creation).”192 In other words, he presumes intentional action on

God’s part. However, he also argues that in light of the evaluative criteria of modern

science we must be open to making certain changes or modifications regarding our

understanding of the reality of God. He has made the argument that “…this concept of a

God who acts in the world has become increasingly difficult to defend in the face of our

191 Nicholas Saunders, Divine Action & Modern Science (Great Britain: Cambridge University Press, 1996), xvii. 192 Ibid., xii. 92

modern scientific worldview.”193 This may be why some regard Saunders’ range of

concepts of God to be limited as it relates to the Christian tradition. Saunders himself

advocates for an approach that will “…consider our current hesitant models of God’s action

in all their details while simultaneously acknowledging the inherent limitations of our

human perspective and the provisional nature of our models.”194 He maintains that for faith

in God to have coherence in our age, there must be a recognition that both science and

theology make overlapping claims relating to the same reality. Therefore, theological

claims must be made in light of scientific knowledge, and scientific knowledge must be

considered in light of theological understanding.195 This seems to be Saunders’ way of

maintaining that while physics and theology are distinct intellectual disciplines asking

different questions about reality, they can dialog with each other since their claims about

reality interact. His view seems to have some alignment with Polkinghorne’s contention

that the critical realism of natural science has its counterpart in theology, meaning that

the search for knowledge and understanding does not apply solely to the natural scientist,

but also to the theologian.

Saunders specifically criticizes Polkinghorne’s claim that the deterministic equations

of chaos theory emerge from the more complex and open behavior of physical reality.

Polkinghorne argues that:

…strict notions of strange attractors and sensitive dependence derive from the mathematical analysis, but surely, one may expect that concepts close to these will, in an appropriate domain, characterize also the true physical theory (which I regard as giving rise to the deterministic equations of chaos as a downward emergent approximation).196

193 Nicholas T. Saunders, “Does God Cheat at Dice? Divine Action and Quantum Possibilities,” Zygon 35, no. 3, (2000): 518. 194 Saunders, Divine Action & Modern Science, xvii. 195 Ibid., x. 196 John Polkinghorne. “Response to Wesley Wildman’s ‘The Divine Action Project’,” Theology and Science 2, no.2, (2004): 191. 93

According to Saunders this is a “deceptively complex argument” regarding the emergence of determinism out of the indeterminism of physical reality.

Whilst Polkinghorne accepts that mathematical chaos is essentially deterministic, crucially he denies that real-world chaos is like this. At the root of his argument is an implicit distinction that runs through Polkinghorne’s work between deterministic mathematical chaos and indeterministic real-world chaotic-like phenomena.…What Polkinghorne does is to turn the standard notion of ontological emergence on its head, and assert that as one moves down from this true supple indeterministic reality, one eventually meets the deterministic equations of mathematical chaos.197

He points out that that most criticisms of Polkinghorne’s reliance on chaos theory fail to account for the subtleties of his arguments. More particularly, he argues that the critics who focus on the deterministic basis of mathematical chaos fail to appreciate “…that

Polkinghorne’s argument is itself dependent on the very determinism that they hold against him.”198

In making his argument Saunders points out the need to establish an for what is meant when we talk about “determinism.” This is particularly important since he identifies eight different forms of determinism used in philosophical discussions and Wildman mentions that there are myriad subtleties concerning the definition of determinism in philosophical writings. In the context of the discussion of

God’s relationship to the world, including God’s action in the world, Saunders relies on

William James (1842-1910) and Richard Montague (1930-1971) for establishing a workable operational definition of determinism as it is used in natural science. James associated “…determinism with a denial of the ontological status of alternatives and indeterminism with the existence of open possibility,”199 while Montague associated determinism with a particular theory. For his purpose Saunders is specifically interested

197 Saunders, Divine Action & Modern Science, 190-191. 198 Ibid., 196. 199 Ibid., 85. 94 in scientific theory. Therefore, characterizing Montague’s approach Saunders states that

“…a theory is deterministic if for any two such systems: if, at a particular given time, they are in an identical state then, according to the theory, they will at all future times be in identical states to each other.”200 This resonates with Wildman’s definition of physical determinism when he explained that “…given that the world is a particular way at one moment, its unfolding thereafter is fixed and inflexible.”201

These characterizations of what is meant by determinism in natural science generally align with Polkinghorne’s characterization of a deterministic universe as one in which “…full knowledge of the present grants retrodictive knowledge of the past and predictive knowledge of the future (This is the argument of Laplace’s celebrated calculating demon.)”202 Recall that his mention of Laplace’s demon is a reference to the claim by the French mathematician Pierre-Simon de Laplace that if one (the demon) knows the precise data for every atom in the universe then their past and future values can be precisely calculated using the classical mechanical laws of physics. This was the view of mechanistic determinism that was overthrown by the revolution brought about by the development of relativity and quantum theories in the twentieth-century.

Polkinghorne’s characterization of determinism and indeterminism is in accordance with the most widely accepted interpretation of quantum mechanics, the Copenhagen interpretation. Werner Heisenberg was one of the founders of quantum mechanics and introduced what has come to be known as the Heisenberg Uncertainty Principle. He explained that we have the ability to translate results of observations into mathematical

200 Ibid., 86. 201 Wildman, “The Divine Action Project, 1988-2003,” 39. 202 Polkinghorne, Belief in God in an Age of Science, 70. 95 equations. With these equations we use measurements of certain physical quantities to derive the same types of quantities for determining the state of a system at a later time.

Furthermore, the ability to do this is due to directing our attention to a part of the universe that becomes the object of our studies.203 We are only able to examine various parts of the universe, isolated from ourselves and the rest of the universe, in the macroscopic realm. However, this is not possible in the quantum realm. Any object of study in this realm is connected to and affected by every other part of the universe. This, then, introduces the uncertainty inherent in quantum mechanics. While the point here is not to become immersed in a philosophical discussion of determinism and indeterminism, it is important to try to grasp, as best as we can, the intentions of Polkinghorne’s critics in making the type of criticisms that they do with regard to his own approach to the question of God’s relationship to the world, as well as what he is saying about the nature of physical reality.

As stated earlier, Saunders proposes to examine Polkinghorne’s motivations for making the assertion that mathematical chaos theory, based in deterministic laws, models an indeterminate physical reality. Saunders maintains:

Polkinghorne’s assertion that real-world chaos is indeterministic cannot be attacked by reiterating the deterministic basis of chaotic mathematics. A correct basis on which to critique this postulate would be to consider Polkinghorne’s motivations for asserting it including features such as strange attractors and sensitive dependence which, as we have seen, exist as an uneasy fusion of the deterministic consequences of mathematical chaos onto an assertion of indeterminacy.204

To make his point Saunders relies on an example of chaotic behavior in some chemical reactions. He asks whether Polkinghorne’s model of active information input into chaotic

203 Heisenberg, Physics and Philosophy: The Revolution in Modern Science, 45, 52. 204 Saunders, Divine Action & Modern Science, 196. 96 systems can actually apply to physical systems in nature. His answer is that “The chaotic modeling of chemical processes assumes an infinite complexity in the values of the concentrations of the various reactants and this simply cannot be the case in nature.”205

The reason that this cannot be the case with nature, according to Saunders, is that for a given quantity of solvent there is a finite number of atoms for any one species.

Therefore, while some chemical reactions manifest many of the features of chaos, a chaotic model that assumes the existence of an infinite intricacy of attractors (constraints) cannot represent physical reality. Moreover, he maintains that this situation also applies to physical mechanical systems as well. The implication here is that such systems do not have the openness or freedom indicative of indeterminate systems. Furthermore,

Saunders suggests that Polkinghorne’s examples of divine action by way of information input may require a model of quantum chaos.206

In his examination of a variety of approaches to the question of God’s relationship to the world Saunders has come to the conclusion that the current understanding of God’s relationship to the world is not as well developed as many maintain. For him this means that contemporary theology is in a crisis. He argues that all of the current models of

God’s interaction with the world which make detailed use of quantum mechanics and chaos fail. However, in his view this “…should not lead us to a Farrer-like desperation on the causal joint issue.”207 Saunders believes that we can still have confidence that God is active in the physical world. The need he identifies is for theologians to continue to work more to address what he calls “the intractable problems” surrounding the question

205 Ibid., 198. 206 Ibid., 199. 207 Ibid., 215. 97 of God’s relation to the world and to become more scientifically informed in the process.208 Saunders’ mention of a “Farrer-like desperation” above is in reference to

Austin Farrer (1904-1968), Anglican priest and theologian, whose view of God’s interaction with the world has been characterized as “double agency”. Farrer used analogy to consider God’s interaction with the world and argued that “…God’s agency does not strike us at the springing-point of causes but in the finished product.”209 He never claimed to be able to identify the juncture between the human and divine, or

“causal joint.” In his words, “The causal joint (could there be said to be one) between

God’s action and ours is of no concern in the activity of religion; the very idea of it arises simply as a by-product of the analogical imagination,…”210 This lingering mystery was not necessarily problematic for Farrer. Therefore, it is against this view of Farrer that

Saunders expresses confidence of coming to better understand God’s interaction with the world.

Wildman doubts that a correspondence can be shown to exist between mathematical chaos and chaos of physical systems, and Saunders believes that physical systems in nature lack the openness of mathematical chaos. Both of these views derive in large part from the uncertainty with regard to the relationship between quantum theory and the macroscopic chaos theory. Polkinghorne has acknowledged that this uncertainty has remained a major problem plaguing the discussion regarding how our everyday world of sense experiences emerges from its quantum substrate; which is another way of saying that there still remains the mystery of the measurement problem. He offers that it is

208 Ibid., 216. 209 , Faith and Speculation (New York, New York: University Press, 1967), 63. 210 Ibid., 66. 98

embarrassing to admit that for more than eighty years after the discovery of quantum

theory there is no satisfactory answer to this foundational scientific question. In his

words:

Microscopic quantum theory and macroscopic chaos theory are imperfectly reconciled to each other. The extreme sensitivity of chaotic systems means that their behavior soon comes to depend on fine details of circumstance to which the quantum uncertainty principle forbids access. Yet a synthesis of the two theories is frustrated by their mutual incompatibility.211

While frustrating, this unanswered question is not necessarily a reason for pessimism or

despair. The scientist and the theologian each one learns to live with intellectual

uncertainty, and having partial answers to questions long pondered serves as a driving

motivation for further research and investigation for those in both disciplines searching

for answers to these lingering questions.

Recall that Wildman mentions the potential value of a new kind of chaos theory

and that Saunders states that Polkinghorne’s proposal may call for a model of quantum

chaos. It is therefore worth mentioning here that there has been recent work done in the

area of quantum chaos. Robert Brecha, Professor of Physics at the University of Dayton,

has argued that the idea of quantum chaos is not clearly understood. According to him:

As the name implies, quantum chaos could be considered to provide a bridge between the world of quantum mechanics and the world of chaotic behavior in nonlinear dynamical systems. It should be noted, however, that it is difficult even to define precisely what we mean by quantum chaos.212

He points out that correspondence between the quantum and chaotic realms at present is

unclear, and in the view of some even contradictory. However, while some advocate for

a retreat from searching for a correspondence between the quantum and macroscopic

211 John Polkinghorne, Quantum Physics and Theology: An Unexpected Kinship (New Haven Connecticut: Yale University Press, 2007), 69-70. 212 Robert J. Brecha, “Schrödinger’s Cat and Divine Action: Some Comments on the Use Of Quantum Uncertainty to Allow For God’s Action in the World,” Zygon 37, no. 4, (December 2002): 920. 99 realms, Brecha argues that physicists of recent years have made some progress in their experimental search for whether or not there exist signs of quantum chaos. This will be addressed in a little more detail later on as part of the evaluation of Wildman’s and

Saunders’ criticisms of Polkinghorne’s view. As part of Brecha’s argument he describes an experiment aimed at finding signs of what might be regarded quantum chaos:

In the recent experiments, the starting point is a “quantum” system consisting of a sample of atoms confined to a small area by light, which can be arranged such that the atoms slosh back and forth slightly. Interactions are arranged to be analogous to a classical system, that is, the group of atoms is periodically given a “kick” in the form of a pulse of laser light.213

As the experimenters monitored the parameters displaying the signature of chaos in the analogous classical system (the “kicked rotor”, which is a swinging pendulum that is periodically given a push), they observed no sign of chaos in the quantum system.

However, the second portion of the experiment involved coupling this system to its surroundings, making it progressively more classical. Per Brecha, “As they increase the amount of classicalizing interaction, they find that the experimental signature of chaotic behavior gradually begins to appear.”214 Perhaps this line of research will eventually provide the knowledge necessary for determining whether such a thing as quantum chaos actually exists. On the other hand, it may reveal that the subatomic realm is not only closed to us with respect to knowing certain physical quantities, but closed to our being able to establish whether chaotic behavior occurs at the quantum level. Either way it may be that a better focus for theologians and scientists engaged in the discussion of God’s relation to the world would be to direct their focus toward endeavoring to better understand and characterize the link between the quantum and macroscopic realms. This

213 Ibid., 921. 214 Ibid. 100 would be invaluable to the scientific community in seeking to better know how the world around us works, as well as a great help to many theologians engaged in discussing the question of God’s relation to the world.

101

CHAPTER 5

EVALUATING CRITICISMS OF POLKINGHORNE

To evaluate the criticisms of Polkinghorne previously mentioned, it may be best

to begin with foundational questions that underlie his reliance on quantum theory and

chaos theory. Recall that the demise of the view of the universe as a great machine came

about by way of two great revolutions in modern physics - Einstein’s theory of relativity

and quantum theory, which included several founders including Max Planck, P.A.M.

Dirac, Niels Bohr, Werner Heisenberg, and Erwin Schrödinger. These revolutions

revealed that space and time are not the backdrop against which phenomena in the

universe are experienced, but rather are phenomena themselves, and that the universe has

revealed itself to be indeterminate at the level of the subatomic. In other words, physical

reality behaves in such a way as not to allow us to know everything about it. Even if we

could imagine having all knowledge possible about the world around us, our knowledge

would be limited by this same world.

Polkinghorne makes the argument that quantum theory actually played a minor

part in the move from viewing the universe as a great machine that can be known with

certainty to something more mysterious where physical quantities at subatomic levels

cannot be known.

Obviously the clockwork universe could not survive the dissolution of the picturable and predictable into the cloudy and fitful, produced by the advent of quantum theory. But, in fact, that theory has only had a minor part to play in the demise of the mechanical. The seeds of its actual decay lay within Newtonian theory itself.215

215 John Polkinghorne, “The Quantum World” in Physics, Philosophy and Theology, ed. Robert John Russell, William R. Stoeger, S.J. and George V. Coyne (Vatican City State: Vatican Observatory, 2000, Berkeley, California: Center for Theology and the Natural Sciences, 2000), 334. 102

As part of his argument he offers that chaos theory is the “real coup de grace” for a

mechanical view of the world since it operates at the macroscopic level, the level of

Newtonian physics, the level in which we live, operate, and register all of our sense

experiences. His reliance on chaos theory, along with quantum theory, lies at the core of

the criticisms of his view that were characterized earlier.

As Polkinghorne points to the fact that chaotic behavior can arise from

deterministic mathematical equations, the underlying metaphysical question that surfaces

has to do with whether the future is ultimately determined, or that the world is open in

some of its processes.216 Furthermore, this leads to the consideration of whether quantum

theory settles this question once and for all. In other words, observed regularity in the

macroscopic realm can be said to simply come about due to “…the statistical effect of the

law of large numbers, the essentially predictable average of many stochastic events.”217

In addition, since chaotic systems possess the characteristic of extreme sensitivity to their

initial conditions, does this not eventually lead to the need to know the quantum-level

values that ultimately define these conditions? There are those who hold the view that

quantum physics ultimately settles the question of macroscopic openness.

If we begin with the hypothesis that God works at the quantum level, it is not necessary – in fact it is counterproductive – to argue for causal indeterminism at higher levels of organization (excluding the human level) since God’s will is assumed to be exercised by means of the macro-effects of subatomic manipulations.218

For those adhering to this view the fact that macroscopic objects are composed of the

subatomic entities identified in quantum physics means that God’s interaction with the

216 Polkinghorne, Reason and Reality, 39. 217 Ibid., 40. 218 Nancey Murphy, “Divine Action in the Natural Order: Buridan’s Ass and Schrödinger’s Cat” in Chaos and Complexity: Scientific Perspectives on Divine Action, ed. Robert John Russell, Nancey Murphy, and Arthur R. Peacocke (Vatican City State: Vatican Observatory, 1995; Berkeley, California: Center for Theology and the Natural Sciences, 1995), 327. 103 world involves bottom-up causation. This includes the argument that “God governs each event at the quantum level in a way that respects the ‘natural rights’ of the entities involved.”219 At first glance quantum physics’ answers to these foundational questions seem to establish the openness to physical systems from a bottom-up view.

However, Polkinghorne identifies three complications that come to light based on foundational questions. These complications are presented here to emphasize that assuming quantum theory definitively answers the question of whether the future is open or determined is somewhat naïve, or at least premature, in that it does not take into account the issues with which physicists have been dealing since the advent of quantum theory. As a particle physicist Polkinghorne has long wrestled with these open questions.

Accordingly, he has kept these in mind in arguing for his view of God’s relationship to the world. The first complication has to do with the character of quantum physics, and in particular whether chaotic behavior exists at the quantum level.220 Since the Schrödinger equation lies at the heart of the Copenhagen interpretation of quantum theory, the mathematical form of this question is whether or not the Schrödinger equation models chaotic behavior. Earlier it was shown that it has been demonstrated that some systems exhibiting chaotic behavior at the macroscopic level are not observed to be chaotic at the quantum level. So there is a lingering question as to whether quantum chaology even exists.

The second complication also relates to the nature of quantum mechanics.221

More specifically it relates to the fact that while the Copenhagen interpretation is the

219 Ibid., 343. 220 Polkinghorne, Reason and Reality, 41. 221 Ibid., 40-41. 104

prevailing view of quantum behavior, it is not the only view. Recall that in the

Copenhagen interpretation the time-evolution of a quantum system means it exists in a

superposition of states (all the states that satisfy the Schrödinger equation) until an

observation or measurement occurs. Up to the point of observation this process is

completely deterministic. However, the instant an observation is made the system

assumes a specific single state, registering a result at the macroscopic level. How that

result is chosen or comes about from the quantum event that exists in all possible states at

the same time is completely non-deterministic. This is the measurement problem referred

to earlier. Contesting views have been offered up in an attempt to solve this

measurement problem. Two of these have risen to some prominence among various

physicists. One of these is the “hidden variables” view proposed by David Bohm (1917-

1992). Simply stated this view argues that when it comes to quantum behavior “…the

apparent unpredictability might just be due to the operation of undisclosed effects –

hidden variables they are called.”222 In other words, in this view quantum entities are

actually determined, but by mechanisms of which we are unaware at present. Another

view that has competed with the Copenhagen interpretation is the “many-worlds” view,

proposed in 1957 by Hugh Everett III, who inferred it from the mathematics of quantum

mechanics, but did not live to see it given the respect that some physicists give it today.223

This view essentially argues that rather than all possible outcomes of a quantum system

existing simultaneously in a superposition of states, whereby only one outcome is

realized when a measurement is taken, every possible outcome is realized, with each

222 John Polkinghorne, The Quantum World (Princeton, New Jersey: Princeton University Press, 1984), 56 223 Peter Byrne, “The Many Worlds of Hugh Everett,” Scientific American, October, 2008 https://www.scientificamerican.com/article/hugh-everett-biography/ 105 being realized in a separate universe. Therefore, the question lingers concerning whether a more adequate explanation of quantum behavior exists that may dispense with indeterminism.

The third complication directly relates to the measurement problem.224 In other words, how is measurement to be characterized? The fundamental question regarding what constitutes a measurement is one with which physicists have been wrestling since the advent of quantum theory. Werner Heisenberg characterized measurement as a permanent result that is recorded in the macroscopic realm. He offered that “If we want to describe what happens in an atomic event, we have to realize that the word ‘happens’ can apply only to the observation, not to that state of affairs between two observations.”225 In other words, it does not make sense to think of the specifics of quantum behavior apart from an observation, or measurement. Niels Bohr offered that a measurement is the interaction of a macroscopic apparatus with the quantum realm, whereby “…no sharp separation can be made between an independent behavior of the objects and their interaction with the measuring instruments which define the reference frame.”226 In other words, any instrument of measurement is composed of microscopic components, which operate according to the rules of quantum mechanics. One cannot distinguish completely between the macroscopic device and its microscopic constituents.

Eugene Wigner proposed that the act of measurement ultimately requires the intervention of human consciousness, which plays a determining role. According to him:

…the impression which one gains at an interaction, called also the result of an observation, modifies the of the system. The modified wave function is, furthermore, in general unpredictable before the impression gained at the interaction has

224 Polkinghorne, Reason and Reality, 41. 225 Heisenberg, Physics and Philosophy: The revolution in Modern Science, 54. 226 Neils Bohr, Atomic Physics and Human Knowledge (New York: Dover Publications, 2010), 52. 106

entered our consciousness: it is the entering of an impression into our consciousness which alters the wave function because it modifies our appraisal of the probabilities for different impressions which we expect to receive in the future. It is at this point that the consciousness enters the theory unavoidably and unalterably.227

He offered that the influence of consciousness is an example of a system possessing non- linearity which breaks the linear superposition of quantum mechanics. Recall that

“…quantum theory not only has states in which a particle is located ‘here’ or is located

‘there’, but it also has states which are a mixture of these two.”228 This is what is meant by the “superposition principle” in quantum theory. Therefore, Wigner was referring to the human observer’s interruption of the deterministic and continuous of the state of the quantum system. When this occurs, the system’s state becomes one that yields a specific measurement result of the of interest. Polkinghorne seems to align himself with Bohr and Heisenberg when he says that measurement refers

“…to the irreversible macroscopic registration of a property that is assigned to the microscopic quantum entity involved.”229

These foundational questions remain part of the framework for the discussion of

God’s relation to the world. Efforts to arrive at more adequate answers to these questions will continue to play a significant role in the discussion. More particularly, they will affect whether and how those engaged in this discussion might need to restructure their views. For the present these questions continue to present complications with which we have to live as we endeavor to grasp Polkinghorne’s approach and the criticisms of it that have been made.

227 Eugene Wigner, Symmetries and Reflections: Scientific Essays (Cambridge: MIT Press, 1970), 175-176. 228 Polkinghorne, Reason and Reality, 86. 229 Polkinghorne, “Physical Process, Quantum Events, and Divine Agency,” 186. 107

The criticisms of Polkinghorne made by both Saunders and Wildman ultimately

center on the question of determinism and indeterminism in nature. More specifically,

their criticisms both focus on the notion that observed chaotic behavior when modeling

physical systems using mathematics does not mean that the world at the macroscopic

level displays indeterminacy and is therefore genuinely open. Wildman argues that no

correspondence principle exists whereby mathematical chaos tells us something about

indeterminacy in nature and Saunders argues that since systems in nature do not possess

the infinite complexity of mathematical chaos then nature cannot display true openness at

the macroscopic level. Essentially, then, both Wildman and Saunders argue that there is

no real alignment between mathematical chaos and a universe that demonstrates openness

at the macroscopic level. Earlier we considered the operational definition of determinism

used by Wildman and Saunders in characterizing their criticisms. To evaluate these

criticisms, it will now help to revisit determinism in greater detail.

Another participant in the conversation regarding God’s relation to the world

provides some insight regarding determinism as it is used in natural science. Dr. William

R. Stoeger, S. J (1943-2014), was ordained a Jesuit priest in 1972 and subsequently

earned his Ph.D. in astrophysics from Cambridge University in 1976, where one of his

classmates was Steven Hawking (1942-2018). From 1979 until his death in 2014 he

served as a staff astrophysicist at the Vatican Observatory in Tucson, Arizona. In the

1980s he joined the board of directors of the Center for Theology and Natural Science

(CTNS).230 The purpose here is not to examine Stoeger’s view of God’s relation to the

world in detail so as to be able to compare and contrast his and Polkinghorne’s views.

230 Robert John Russell, “William R. Stoeger, S.J. (1943–2014): Physicist, Cosmologist, Friend, and Leader in Theology and Science,” Theology and Science 12, no.4, (2014): 293-295. 108

However, a brief examination of Stoeger’s view involves his insights with regard to the

nature of determinism that may be useful in evaluating criticisms of Polkinghorne

mentioned earlier. Stoeger, like Polkinghorne, approaches the question of God’s relation

to the world both as an ordained clergyman as well as a professional scientist. He

confesses a view of God as:

Trinity of persons – with all that that implies within the Christian tradition – then all divine action however, impersonal it may seem, in its consequences or manifestations, must be seen in terms of the personal, of personal relationships, and of the preconditions of the emergence of the personal within the universe. This is certainly true from the standpoint of our faith and knowledge which we have based on divine revelation. However, it is far from clear simply from the standpoint of the physical and the other natural sciences, even though there are indications that point in that direction (e.g., the coincidences which point towards an “anthropic principle”, however, vacuous the actual logic of those arguments may be without the presupposition of God’s existence).231

He understands God’s self-communication as “…an act of efficient causality (but not

limited to efficient causality) producing of things not God in being.”232 These things are

utterly dependent on God, who is distinct from creation. God’s creative action does not

complete God; still God is immanent within creation. Similar to Polkinghorne, he

articulates the belief that God has been, and continues to be, actively present and

involved in our daily lives as creator, sustainer and interactor.

Stoeger also emphasizes that the disciplines of science and theology can dialog

with each other. While the languages of these two disciplines are different, and they

address different questions, this does not mean that they are totally isolated from each

other. Instead he argues that “They continue to be in dynamic interaction in our common

231 William R. Stoeger S.J., “Describing God’s Action in the World in Light of Scientific Knowledge of Reality” in Chaos and Complexity: Scientific Perspectives on Divine Action, ed. Robert John Russell, Nancey Murphy, and Arthur R. Peacocke (Vatican City State: Vatican Observatory, 1995; Berkeley, California: Center for Theology and the Natural Sciences, 1995), 259. 232 William R. Stoeger, S. J. “Conceiving Divine Action in a Dynamic Universe” in Scientific Perspectives on Divine Action: Twenty Years of Challenge and Progress. ed. Robert John Russell, Nancey Murphy, and William R. Stoeger, S.J. (Notre Dame, In.: University of Notre Dame Press, 2008), 230. 109 cultural and academic fields.”233 Accordingly, his stated aim is to articulate a view of

God’s relation to the world in a manner that remains faithful to Christian sources of revelation and that is consistent with what the natural sciences teach us about the world around us. Therefore, he endeavors to state his view in light of the description of the general character of God as known through contemporary Christian belief and theology and in light of descriptions of the general character of the world as known through contemporary science.234

Regarding revelation and our reflection on it Stoeger offers that God answers the foundational question “Why there is something rather than nothing?”235 Like

Polkinghorne, he holds that God interacts with the world, and offers that God’s motivation for creation and interaction with it springs from God’s goodness and desire to share that goodness and divine love, both for God’s self as well as for all that God creates and holds in existence. He maintains that creation is good and an expression of God’s goodness and love. This means that creation itself reflects something of God. This leads to his conclusion that God reveals something of God’s self though creation, and that

God’s most particular revelation comes by way of persons and personal relationships involving love and forgiveness. Accordingly, we experience God as personal, who gives love, forgiveness, and reconciliation. Stoeger maintains that interpersonal relationships are of paramount importance to God, which is true of God as Trinity and God’s external relationships.236 Therefore, Polkinghorne, a particle physicist and Anglican priest, and

233 Stoeger, “Describing God’s Action in the World in Light of Scientific Knowledge of Reality,” 240. 234 Ibid., 239. 235 Ibid., 245. 236 Ibid., 245. 110

Stoeger, astrophysicist and Catholic priest, both focus on the relational nature of God in proposing their respective views of God’s relation to the world.

Another similarity between Polkinghorne and Stoeger is that both endeavor to present a view of God’s relation to the world that is consistent with what modern physics reveals to us about physical reality. Therefore, with regard to the world around us as known through contemporary natural science Stoeger offers that every level within nature manifests self-organizing and self-ordering principles.

…at every level there are self-ordering and self-organizing principles and processes within nature itself, which can adequately describe and account for (at the level of science) its detailed evolution and behavior, the emergence of novelty, possibly even of consciousness, the inter-relationships between systems and levels, and even the various laws of nature themselves.237

In other words, nature displays novelty, which seems to align with Polkinghorne’s discussion of “free process” and the emergence of novelty in nature. Polkinghorne argues that “…the interlacing of order and disorder is precisely what seems to be needed for the creative emergence of novelty.”238 As an example of an experiment for demonstrating such behavior he describes a configuration to simulate dissipative systems, or systems which are open and exchange energy and/or matter with their surrounding environment. This configuration consists of a large array of electric light bulbs where each individual bulb possesses two possible states – it can either be “on” or “off”. Rules built into the experiment determine the system’s state at any subsequent time based on its present state. The system is configured such that each bulb is correlated with two other bulbs elsewhere in the array. The rules of the system specify how the state of the two

237 Stoeger, “Describing God’s Action in the World in Light of Scientific Knowledge of Reality,” 242. 238 Polkinghorne, Exploring Reality (New Haven, London: Yale University Press, 2005), 27. 111 bulbs determine the individual bulb’s next state. The array is started operating in a state of random illumination, with some bulbs on and some off.

It is then left to develop according to the rules. One might have expected that generally nothing very interesting would happen and that the array would just twinkle away haphazardly for as long as it was left to do so. In fact this is not at all the case. If there are 10,000 light bulbs in the array, the number of possible states of illumination possible in principle is 210,000, or about 103000, an absolutely huge number. Yet, the system soon settles down to cycling through only about 100 states! This phenomenon represents the spontaneous generation of an astonishing degree of order. Once again, more seems to be different in a highly non-trivial matter. Similar spontaneous generation of complex order is also found to occur in systems of cellular automata.239

In his view the unexpected novel behavior demonstrated here illustrates that “In those cases where chaos is generated through a cascade of bifurcating possibilities, there is a remarkable universal pattern in the way this happens, characterized by a new mathematical constant of fundamental significance.”240

An emphasis Stoeger makes is that “The gaps in scientific knowledge have not all been filled, but they are gradually being filled by new discoveries. And it has become clear that appealing to divine intervention is not an acceptable means for doing so.”241

This aligns well with Polkinghorne’s criticism of a “God-of-the-gaps” theology whereby people appeal to God in an effort to fill the epistemological gaps in human knowledge that science may subsequently fill. Stoeger holds that “The laws of nature and nature itself constrain but under determine what develops or occurs.”242 The key examples he cites to emphasize this point are quantum indeterminacy, chaotic unpredictability, and freedom in the establishment of initial and boundary conditions for various physical systems and processes.243

239 Ibid., 29. 240 Ibid., 28. 241 Stoeger, “Describing God’s Action in the World in Light of Scientific Knowledge of Reality,”243. 242 Ibid. 243Ibid., 244. 112

His characterization of determinism is that “A system is deterministic if its state or configuration at any time automatically leads to (determines) a single definite state or configuration at a later time.”244 In other words, once the state of the system is established at a specific time, then inevitably the system will be in a definite state at any given subsequent time, if there are no external influences that work to change the system’s parameters. Given that the system parameters are precisely known at one time, a physicist will use the mathematical equations governing the system to predict the precise state of the system at a later time. This aligns with Polkinghorne’s characterization of a deterministic universe as one in which complete knowledge of the present state of affairs yields the ability to develop complete knowledge of the past as well as the future states of affairs.245

Given this understanding Stoeger argues that determinism in nature is “system specific”, meaning “…we can only apply the term to a definite well defined system, which must be isolated from other systems. If it is not isolated, then we must examine the smallest larger completely isolated system of which the original system is a part – to see if that system is deterministic.”246

Furthermore, he points out that in most cases determinism is also “level specific.”

That is to say just because a system is found to be deterministic on one level of organization does not imply “…either that the less complex systems on more fundamental levels which constitute the system on the given level, or the more complex systems on higher levels, which are constituted by the systems on the given level, are

244 Stoeger, “Conceiving Divine Action in a Dynamic Universe,” 246. 245 Polkinghorne, Belief in God in an Age of Science, 70. 246 Stoeger, “Conceiving Divine Action in a Dynamic Universe,” 246. 113 deterministic.”247 Therefore, this makes it extremely difficult, if not impossible, in practice to know whether a particular physical system is deterministic. This does not mean that there are no idealized cases in which it is relatively easy to demonstrate that a system is deterministic. However, it is much more challenging to test whether a system in nature is isolated, or whether the larger system of which it is a part is isolated in such a way to allow us to check deterministic character. It would seem, then, that for Saunders to claim that real-world systems do not actually display openness he will need to be more specific as to which system or systems he is referring, as well as include more information about those systems such as all of their associated levels and whether the systems can be isolated from other systems. More information than the fact that a particular physical system is not infinite is needed in order to argue that a system is determinate. To be fair when Saunders used an example of a chemical system in his criticism of Polkinghorne he was arguing that the system cannot manifest real openness because it does not possess infinite intricacy. However, the implication in his argument seems to be that the system, or any real-world system, is ultimately determinate because it lacks the genuine openness of indeterminate systems.

When one considers the universe or any number of its subsystems as they really exist and not merely how they are modeled mathematically, being able to test whether they are deterministic is replete with difficulty. Heisenberg pointed out that the quantum substrate of physical reality induces indeterminism into any registration of a result at the macroscopic level. This is part of the measurement problem whereby the measuring device introduces a new element of uncertainty since “…it contains all the uncertainties

247 Ibid. 114 concerning the microscopic structure of the device…”248 However, aside from the measurement problem there are myriad levels of complexity intertwined in natural systems in such ways as to make it virtually impossible to isolate a macroscopic system for study in order to establish whether or not it is deterministic. In addition to this is the point that “…determinism on any level does not, as we have already seen, trickle up or down levels.”249 The argument here is that determinism cannot be applied to nature as a whole, even though there may be limited cases where it is observed in given levels of certain subsystems. In other words, “…strict determinism is not a characteristic of nature as a whole.”250

This characteristic of determinism as level-specific could potentially apply to what Polkinghorne is arguing with regard to determinism’s manifestation at some levels and not others:

To address the issue bluntly: if apparently open behavior is associated with underlying apparently deterministic equations, which is taken to have the greater ontological seriousness – the behavior or the equations? Which is the approximation and which is the reality? It is conceivable that apparent determinism emerges at some lower levels without its being a characteristic of reality overall.251

According to Polkinghorne “…a simple and perfectly deterministic equation can produce behavior which is random to the point of unpredictability.”252 In addition, such systems have the characteristic that unless one knows the system’s initial conditions with unlimited accuracy, predicting its behavior into the future can only be done in a small way with any confidence. This is what he means by referring to these physical systems as intrinsically open. Moreover, Polkinghorne is not arguing that chaotic systems in

248 Heisenberg, Physics and Philosophy: The Revolution in Modern Science, 53. 249 Stoeger, “Conceiving Divine Action in a Dynamic Universe,” 247. 250 Ibid., 248. 251 Polkinghorne, Reason and Reality, 41. 252 Ibid. 36. 115

nature have an infinite range of outcomes but rather that “They do not wander all over the

place but their motions home in on the continual and haphazard exploration of a limited

range of possibilities (called a strange attractor).”253 Mathematically a strange attractor

has infinite complexity, giving it the quality of zoom symmetry mentioned earlier. The

strange attractor is created by repeated mathematical iterations, without end, of the same

pattern. Therefore, when it is graphed it produces a pattern (fractal) with the

characteristic that however much one zooms onto the pattern one will encounter the same

pattern with which one started. Mathematically the trajectories which constitute the

pattern infinitely fold in on themselves without ever crossing. The pattern itself,

however, is the graphical result of a limited set of solutions to the deterministic equations

that produce them. Therefore, while chaotic physical systems may not possess infinite

complexity, they possess limited sets of possible behavior that reveal “…an orderly

disorder in their behavior.”254 Using the coastline of Britain as an example, the argument

has been made for a correlation between mathematical chaos and real-world systems.

As complicated as the relation between chaos-models and real-world dynamical systems may be, it is true that as science progresses it becomes more and more clear that there be a structural similarity between chaos in mathematics and some real-world dynamical behavior. The similarity might be found in the dynamics. Perhaps the coastline of Britain is not literally fractal, but there are some (intuitive) similarities which suggest that the underlying process might be comparable.255

The key point for Polkinghorne here is not that chaotic systems in nature have infinite

symmetry, but that they exhibit the unpredictability characteristic of chaotic behavior,

and therefore exhibit a correlation to mathematical chaos. Further, it is his contention

253 Ibid., 36. 254 Ibid. 255 Taede A. Smedes “Chaos: Where Science and Religion Meet: A Critical Evaluation of the Use of Chaos Theory in Theology,” in Studies in Science and Theology 8. Yearbook of the European Society for the Study of Science and Theology 2001-2001, edited by N. H. Gregersen, U. Görman, H. Meisinger (Aarhus, Denmark: University of Aarhus, 2002) 292. 116

that the deterministic equations of physics that are formed in the minds of physicists are

approximations derived from a more complex, open reality. Effectively this means that

such systems evidence indeterminacy and genuine openness. It seems plausible,

therefore, that given the modeling of quantum behavior using deterministic mathematics

(quantum theory), openness in the macroscopic realm might also be modeled by

deterministic mathematics (chaos theory). As the language of science, math is the

scientist’s way to model physical reality. Mathematical models in science, then, can

present us with adequate descriptions of physical reality. However, this does not mean

that there is complete alignment between every aspect of the mathematics and physical

reality.

In their criticisms of Polkinghorne’s use of chaos theory both Wildman and

Saunders argue that there is no real alignment between mathematical chaos theory and

the real world. This seems to come close to arguing that all parts of the universe at the

macroscopic level could theoretically be determined, and that it is an epistemological

openness for which Polkinghorne argues, not an ontological one. The problem of

applying strict determinism to the universe as a whole is based “…on whether infinite

knowledge is a practical feasibility.”256 Likely most scientists are convinced that in order

to precisely know the initial conditions of systems in the real-world infinite knowledge is

needed and that this is beyond human capability. Polkinghorne points to this.

The physical world involves processes taking place on all scales, with the truly microscopic and truly macroscopic being endpoints of a continuum. Since there is a single physical world, and not two disjoint worlds of classical and quantum physics, we need a single integrated account of the universe’s process. If we are to talk properly of

256 Taede A. Smedes, “Is Our Universe Deterministic? Some Philosophical and Theological Reflections on an Elusive Topic,” Zygon 38, no. 4, (December 2003): 967. 117

what is happening, the discourse must be able to refer to all levels of complex physical reality. We do not currently possess such an ability.257

Many phenomena in the universe exhibit behavior that is similar to mathematical models

of chaotic systems, which have been characterized as idealistic approximations of these

chaotic physical phenomena. In principle mathematical chaos (chaos theory) is

deterministic in that every state of a system derives from the former state. Therefore,

“Iterating a chaotic equation is a highly deterministic process.”258 However, with regard

to chaotic physical systems in the real world, things become dramatically different.

Computers, for example, are not adequate for determining the detailed history of such

systems. No one can precisely know the initial conditions of these systems, nor precisely

know and predict the ways in which their associated causal chains interact. In order to

obtain precise knowledge of the initial conditions of a physical system or event one

would need instantaneous and complete knowledge of the parameters of every atom in

the universe, and of every law by which nature operates. The universe does not allow us

the access to the data needed to predict the exact outcome(s) of or to know every

contributing cause of an event in nature. In other words, in the real world indeterminate

behavior is observed at the macroscopic and quantum levels.259

Polkinghorne makes the argument that physical chaotic systems, “…are

intrinsically unpredictable and unisolable in character.”260 This leads him to “…propose

that this should lead us to the metaphysical conjecture that these epistemological

properties signal that ontologically much of the physical world is open and integrated in

257 Polkinghorne, “Physical Process, Quantum Events, and Divine Agency,” 186-187. 258 Ibid. 259 Smedes, “Is Our Universe Deterministic? Some Philosophical and Theological Reflections on an Elusive Topic”, 967-968. 260 Polkinghorne, “The Metaphysics of Divine Action,” 153. 118 character.”261 By using the term “open” here he means those principles that determine energy exchanges among the constituent parts of these systems (bottom-up causality) do not exhaustively determine the future behavior of the systems. In other words, there must be room for other causal principles to operate. In addition, by characterizing these systems as “integrated”, he means that the other causal principles have a holistic character (top-down causality).262 He is not applying indeterminism or determinism to the universe as a whole, but rather arguing that there are parts of the universe at the macroscopic level that are open even though limited cases exist where deterministic behavior is observed. He has made sure to emphasize that his focus is on those parts of the world that are open and therefore leave room for the operation of causal principles other than those inherent in the microscopic constituent parts of systems and processes.

In addition to arguing against any correlation between mathematical chaos and the chaos of physical systems, Wildman challenges Polkinghorne’s notion that God’s interaction with the world does not add energy to physical systems, in violation of the law of conservation of matter and energy. Wildman refers to this as “zero-energy action.” In his discussion of zero-energy action it is sometimes not clear whether Wildman is referring to adding energy to a system, or whether he means that there exist energy level differences between potential behaviors of a chaotic physical system. This may become important since it seems that chaotic physical systems possess potential patterns of behaviors resulting from infinitesimal changes to initial conditions, which can include patterns of differing levels of energy. Polkinghorne’s specific argument is that God’s interaction with the world by way of the initial conditions of chaotic physical systems

261 Ibid. 262 Ibid. 119

involves the input of information rather than adding energy (creating energy).

Furthermore, he argues that this active information input results in the top-down choice

of one pattern of behavior among all of the other potential possible behaviors. This

would then mean that no is violated due to God’s interaction with the world

because active information input does not involve the addition of energy to the universe.

As indicated earlier, research has been ongoing to determine whether quantum

chaos might exist and in turn provide a bridge between classical chaos and the quantum

realm. Recent research has been done to address the question of how classical chaos

might emerge from a quantum system.263 Researchers used a collection of cesium-133

atoms, laser light, and magnetic fields to simulate a classical system that “…can be

thought of as a spinning top, in which chaos is induced by a sequence of periodic driving

kicks.”264 This is an experimental configuration somewhat similar to the experimental

setup described by Brecha earlier. These researchers were also able to develop a method

for switching the system back and forth between non-chaotic and chaotic behavior by

controlling atomic spin and orbits. The experimental configuration allowed researchers

to create a non-equilibrium condition in the system in that the system evolves with time,

and in which they could look for fingerprints of chaos in a process called

“thermalization”, which is defined as follows:

The key chaos-related question addressed by the author is whether the quantum system thermalizes, e.g., relaxes to a state in which the expectation values no longer change and which is distributed more or less evenly over the allowed energy manifold.265

263 Eva-Maria Graefe, “Quantum Chaos on Display,” Physics 6, no. 9, (January 22, 2013): 1-3, http://link.aps.org/doi/10.1103/Physics.6.9. 264 Ibid., 1. 265 Ibid., 2. 120

For a measurement taken multiple times on identical systems, the “expectation value” is

defined as the average of all possible outcomes of the measurement as weighted by their

likelihood.266 What researchers found was that when the classical counterpart is chaotic

the quantum system evidences chaotic behavior in that it also thermalizes. Furthermore,

they found that similar to classical systems’ motion along closed, periodic paths under

specific initial conditions, quantum systems also behave in a similar manner leaving

behind evidence (what they term “quantum scars”) that can be observed.

A key result of this research is that scientific investigation might potentially show

evidence for quantum chaos. Therefore, calls for investigation of quantum chaos, along

with the development of theories of quantum chaology, may eventually find some footing

in such experimental research. Second, since this recent research has revealed quantum

expectation values that do not change across myriad forms of energy, this may have the

potential to contribute to a better scientific understanding of energy conservation at the

level of quantum systems exhibiting chaotic behavior. If physics develops a better grasp

of energy conservation at the quantum level, this could potentially enable Polkinghorne

to better express how the law of conservation of energy is accounted for in his view of

information input and top-down causality. Third, such research has the potential to lead

to a better grasp of the emergence of macroscopic chaos from quantum behavior.

Therefore, it is possible that this could have the effect of influencing Polkinghorne, and

other participants in the conversation of God’s relation to the world, to modify their

views in light of a more adequate grasp of the coupling between the quantum and

266 Robert Eisberg and Robert Resnik, Quantum Physics, of Atoms, Molecules, Solids, Nuclei, and Particles (New York, New York: John Wiley and Sons. 1974), 154.

121 macroscopic levels of nature. Moreover, it might possibly shed light on the long- discussed measurement problem, which ultimately seems to reside at the core of the criticisms made against Polkinghorne’s reliance on quantum theory along with chaos theory. Perhaps a better grasp of the indeterminate world of quantum events will shed new light on the discussion of God’s relation to the world.

According to the Copenhagen interpretation of quantum physics the level of quantum events is fuzzy and indeterminate and only yields definite results by way of an act of measurement. Making a measurement of a quantum system effectively means selecting a definite outcome, which occurs by way of the experimental set up chosen for making the measurement. However, rather than involving the addition of energy to the universe, the measurement act involves the utilization of energy already built into the universe. Therefore, the top-down choice of certain initial conditions in macroscopic chaotic systems (active-information input) for which Polkinghorne argues in his view of

God’s interaction with the world must eventually relate in some way to the selection of outcomes at the quantum level. With the mystery still inherent in the quantum level of nature and the wealth of questions with which physicists are still wrestling regarding the measurement problem, it seems premature for critics of Polkinghorne to claim that active-information input cannot occur without creating energy, which would mean a violation of a fundamental scientific law. Polkinghorne himself acknowledges our continuing need for a more adequate understanding of the measurement problem.

Heisenberg encouraged the thought of the wavefunction as being the carrier of potentiality (the superposition of a variety of outcomes) in contrast to its representing the instantiation of a definite state of affairs. Our difficulty in feeling quite at ease with this notion stems from our ignorance of how that general potentiality becomes a particular actuality as a result of observation – in other words, how the quantum world can be understood to interlock with the everyday world, without our having to embrace the crude and unsatisfactory dualism of the original Copenhagen interpretation. We return once

122

more to the central quantum conundrum, the measurement problem. Until we have agreement on how to deal with that issue we shall remain in perplexity about detailed questions of quantum ontology.267

267 Polkinghorne, Reason and Reality, 97. 123

CHAPTER 6

CONCLUDING THOUGHTS

From the first days of discovery regarding how physical reality behaves at the subatomic level along with the subsequent development of modern quantum theory, a foundational question that has essentially remained unanswered has to do with how it occurs “...that the Newtonian world of everyday experience emerges from the fitful quantum substrate?”268 Polkinghorne pointed out that even after more than ninety years physicists have not arrived at a comprehensive answer to this question that has universal agreement. Accordingly, this question has been at the core of alternate proposals to the

Copenhagen interpretation as well as subsequent research in the hope of being able to better characterize the quantum world described by the Copenhagen interpretation. Since the key criticisms of Polkinghorne reviewed here ultimately have some connection with the relation between the quantum and macroscopic levels of physical reality, then it seems that a more productive approach to this aspect of the discussion of God’s relation to the world would be to focus greater attention on efforts to better understand the relationship between these levels of physical reality. While Saunders and Wildman both allude to this effort, it is not really the main focus in their discussions of their criticisms of Polkinghorne’s reliance on chaos theory. As a mathematical theory, chaos theory may not be directly related to quantum theory, but it has an indirect relation in that it must eventually involve the consideration of how the infinitesimal details of initial conditions of certain physical systems arise from the quantum substrate of physical reality. When

268 Polkinghorne, Quantum Physics and Theology, 69. 124

Polkinghorne’s critics make the claim that real-world chaos does not manifest genuine openness they seem to be skipping a more foundational step, namely the acknowledgement that macroscopic results ultimately come about from quantum potentialities. This step seems to be essential in order to begin to draw more informed conclusions about whether or not macroscopic physical reality is open in some instances.

Current knowledge of the relationship between the quantum and macroscopic levels allows limited dialog with regard to whether or not genuine openness exists at the macroscopic level by way of physical chaotic systems. Therefore, it would seem that more complete and informed dialog on this issue could possibly come by way of greater understanding of the quantum – macroscopic relationship. This does not constitute an argument that a definitive answer to the measurement problem can be had. The

Copenhagen interpretation of quantum physics tells us that the measurement problem cannot be solved because it is part of the structure of physical reality. It is referred to as a problem in that human curiosity wants to find a way to understand it, to explain it, but nature reveals that it is something with which we must live. Werner Heisenberg made the point that answers to scientific questions as well as solutions to scientific problems will finally be decided by nature itself, rather than a group of scientists.269 Accordingly he alluded to the fact that nature places a limit on human probing into the workings of the subatomic level. In his words:

Our scientific work in physics consists in asking questions about nature in the language that we possess and trying to get an answer from experiment by the means that are at our disposal. In this way quantum theory reminds us, as Bohr has put it, of the old wisdom that when searching for harmony in life one must never forget that in the drama of existence we are ourselves both players and spectators.270

269 Heisenberg, Physics and Philosophy: The Revolution in Modern Science, 193-194. 270 Ibid., 58. 125

Therefore, the physicist and theologian each are both researcher as well as part of the experiment.

With the advent of quantum physics no longer can the object of interest be isolated from the rest of the world. Through time the scientist and theologian have each learned to live with the difficult and unanswered (or unanswerable) questions. Striving to better understand how the world of quarks and leptons relates to the world of pendulums and rockets still seems to be the best approach in that it is the one that most directly concerns the problems that lie at the heart of the question of the causal joint of God’s interaction with the world. Given the issues on which the critics of Polkinghorne disagree the hope is that this approach will help the participants in the discussion make some progress, rather than to remain at an impasse. Again, this is not to say that anyone can solve the measurement problem; this is not an option by virtue of the way physical reality is structured and behaves. Were there no measurement problem it would also mean that there is no such thing as ontological openness at the quantum level of nature and eventually at the macroscopic level of nature. However, while nature is structured in such a way as to prevent access to certain information, this does not prevent the possibility of gaining a better understanding with regard to some aspects of the relationship between the quantum and the macroscopic. It seems that any better grasp of this relationship would potentially enable more fruitful discussion between those concerned with the endeavor to discuss causal action of God in God’s relationship to the world.

As a high energy particle physicist, Polkinghorne acknowledges the need to better understand the relationship between the quantum world and the macroscopic Newtonian

126 world, which in part involves understanding whether or not chaotic behavior can be observed at the quantum level. As one who relies on chaos theory for articulating his view of God’s interaction with the world, he acknowledges this complex and frustrating question.

The actual difficulty is that macroscopic dynamics cannot possibly be the whole story, since quantum theory clearly operates at microscopic levels. The fractal geometry assigned to strange attractors must therefore be modified on some scale. This implies that the zero energy difference between paths through the attractor, appealed to in support of introducing the notion of pure “active information” cannot hold absolutely, but must be modified, at least at the level of Heisenberg energy uncertainties. Once again we run afoul of the unresolved problems of quantum chaology in any attempt to appeal to a “hybrid scheme.”271

His approach, as well as all of the other approaches relating to understanding God’s interaction with the world, are what he calls “zero-order approximations,” meaning that they are crude starting points from which better developments can possibly be made.

Rather than looking to quantum physics as his starting point Polkinghorne has chosen to begin with chaos theory since he is “…supposing that its direct connection with macroscopic phenomena, where human and divine agency are both expected to manifest themselves, makes it the more promising zero-order approximation.”272 Chaotic systems also have a connection to the quantum level since they eventually provide a means for the enhancement of quantum events by way of making them, or more properly showing them, to be an integral part of these macroscopic systems that manifest extreme sensitivity to the finest details of their initial conditions.273 It is important to understand that Polkinghorne takes his approach fully aware of the problematic unanswered questions. He therefore advocates for the more fundamental step of endeavoring to better

271 Polkinghorne, “Physical Process, Quantum Events, and Divine Agency,” 189. 272 Ibid., 190. 273 Polkinghorne, Faith, Science & Understanding, 120-121. 127 understand the mutual relationship between quantum systems and macroscopic chaotic systems. At the same time he shares his conviction that macroscopic physical reality displays genuine openness in certain instances, and his confidence that a better grasp of the mutual relationship between the quantum and macroscopic levels will bear this out.

Quantum physics has not only revealed the universe to be indeterminate at the infinitesimal subatomic level of nature, but has also demonstrated that absolute knowledge of physical reality is forever closed to us. That is to say, only partial knowledge of quantum entities is allowed by physical reality. This would seem to imply that fostering further progress in the discussion of the relation between God and the world, for the scientist as well as the theologian, means that knowledge relating to the causal joint of divine interaction will also be closed, hidden to some degree. As stated earlier, Polkinghorne is not appealing to quantum theory as the starting point for his approach, but rather to chaos theory due to its direct connection to macroscopic phenomena. However, obtaining precise knowledge of the initial conditions of a physical event by way of acquiring instantaneous and complete knowledge of the parameters of every atom and in the universe eventually leads to the need for complete knowledge of quantum entities as well as how these entities are enhanced to effect macroscopic outcomes. Quantum entities do not permit such complete knowledge. So it may be inevitable that complete, unrestricted knowledge of the causal joint of God’s interaction with the world will also remain hidden to us. Polkinghorne seems to allude to this in articulating his approach.

There are two basic strategies that might be pursued in the hope of fostering further progress. One is that of bold metaphysical speculation, which takes science into account but is relatively uninhibited in pressing on to grander designs. The alternative, more cautious, approach recognizes that science alone will never provide the key to the nature of agency, yet seeks patient clarification of the complex character of physical processes

128

as being the best basis for the elaboration of further metaphysical ideas. I am of this second school. My immediate hope is for the further progress and clarification that the insights of quantum chaology might have to offer.274

Flexibility resulting from the unpredictable sensitivity of response to initial conditions in complex dynamical systems allows the freedom to become and so provides the freedom for God and humans to act. This is essentially the interplay of chance, expressed in potentiality, and necessity, expressed in scientific law. Polkinghorne refers to the way chance is used here as “tame chance”, by which he is referring to “…the shuffling operations by which what is potential is made actual.”275 So the question remains as to whether or not the macroscopic openness which Polkinghorne is proposing as a necessary mechanism for divine interaction with the world, is also what will eventually prevent complete knowledge of the causal joint for this interaction.

If complete knowledge of the causal joint of God’s interaction with the world is hidden to us, then this does not mean that the discussion of God’s relation to the world is at a dead end. Actually, it would seem to mean just the opposite in that this age-old question has come to be discussed in a new context in light of the scientific achievements of the twentieth century that led to viewing the world around us in a revolutionary way.

If the advent of quantum theory, with its revelation that complete knowledge of quantum entities is hidden to us, has not stopped scientific advancement toward more adequate knowledge of quantum entities and how they work, then neither should theological advancement toward better understanding God’s relation to the world be hindered just because complete knowledge of the causal joint of God’s interaction is hidden to us. It is important to recall that with regard to our knowledge of the way things are Polkinghorne

274 Polkinghorne, “Physical Process, Quantum Events, and Divine Agency,” 190. 275 Polkinghorne, Science and Providence: God’s Interaction with the World, 39. 129 argued that “Of course, such knowledge is to a degree partial and corrigible. Our attainment is verisimilitude, not absolute truth.”276 As a physicist his hope is to achieve a greater understanding of the way the world around us works, and as a person of Christian faith his hope is to work toward a greater understanding of how God interacts with the world.

Puzzling questions which are difficult to answer are part and parcel of scientific effort as well as theological work. As a scientist and theologian Polkinghorne recognizes this as well as the fact that that we live in one world with different aspects that are explored by science and theology.

They are both concerned with exploring, and submitting to, the ways things are. Because of this they are capable of interacting with each other: theology explaining the source of the rational order and structure which science both assumes and confirms in its investigation of the world; science by its study of creation setting conditions of consonance which must be satisfied by any account of the Creator and his activity. Their relationship is not free from puzzles but I have sought to show that no act of mental compartmentalism or dishonest adjustment is required of those who take with equal seriousness the stories told by science and by faith.277

This aligns with his effort to make a series of comparisons between quantum physics and the theological pursuit of understanding the nature of God in the Christian context. For modern science human ability to know something about the physical world convinces us of its reality. This remains the case even when the subatomic level of that reality proves to be elusive, beyond our ability to picture. Such pursuit has trained scientists to think in new ways about concepts that are not intuitive, and the argument has been made that this is analogous to theology’s pursuit for an understanding of the “Unpicturable.”278 Since empirical investigation into how the world around us works has stretched minds and

276 Polkinghorne, Belief in God in an Age of Science, 104. 277 Polkinghorne, One World, 97. 278 Ibid., 47. 130 expanded the notion of what is humanly conceivable, it should therefore be surprising if encounter with the person and work of God will not do the same. Polkinghorne offers that quantum “…theory’s novel insights were only embraced after much struggle, rationally motivated by the need to respond to the stubborn way things are. I believe that theology is engaged in an analogously demanding attempt to do justice to our encounter with the Otherness of the divine nature.”279

Confronting the puzzles inherent in the disciplines of science and theology lies at the core of the quest for better understanding the relationship between God and the world.

Given the new context of the way contemporary science views the world around us, continuing the search for a more adequate understanding of the relationship between God and the world should help foster a better grasp, or at least a better appreciation, of the nature of physical reality as well as help contribute to faith’s search for understanding.

279 Polkinghorne, Reason and Reality, 98. 131

BIBLIOGRAPHY

Abell, George O. Exploration of The Universe. New York, Chicago, Dallas, Atlanta, London, Sydney: Holt, Rinehart, and Winston, 1975.

Bacon, Francis. The New Organon. edited by Fulton H. Anderson. Upper Saddle River, New Jersey: Prentice Hall Publishing Company, 1960.

Barbour, Ian G. Religion and Science: Historical and Contemporary Issues. San Francisco: Harper Collins Publishers, 1997.

Bauckham, Richard. “Jürgen Moltmann” In The Modern Theologians. Edited by David F. Ford, Malden, Ma.: Blackwell Publishing, 1997.

Bohr, Niels. Atomic Physics and Human Knowledge. New York: Dover Publications, 2010.

Bohr, Niels. Structure of the Atom. Nobel Lecture, Royal Swedish Academy of Sciences, Stockholm, Sweden, 1922, https://www.nobelprize.org/uploads/2018/06/bohr-lecture.pdf.

Brecha, Robert J. “Schrödinger’s Cat and Divine Action: Some Comments on the Use of Quantum Uncertainty to Allow for God’s Action in the World.” Zygon 37, no. 4 (December 2002): 909-924.

Brooke, John Hedley. Science and Religion. Cambridge, United Kingdom: Cambridge University Press, 2014.

Brooke, John. “Science and Religion: Lessons from History?” Science 282, no. 5396 (Dec. 11, 1998): 1985-1986. http://science.sciencemag.org/content/282/5396/1985.

Campbell, Lewis, Garnett, William. The Life of James Clerk Maxwell. London: Macmillan and Co., 1882, https://archive.org/details/lifeofjamesclerk00camprich/page/404/mode/2up.

Durant, Will and Ariel. The Story of Civilization: The Age of Reason Begins. New York, New York: MJF Books, 1961.

Einstein, Albert and Leopold Infeld. The Evolution of Physics: From Early Concepts to Relativity and Quanta. New York, London: Simon and Schuster, 1966.

Einstein, Albert. Out of My Later Years. New York, New York: Citadel Press, 1984.

Eisberg, Robert and Resnik, Robert. Quantum Physics, of Atoms, Molecules, Solids, Nuclei, and Particles. New York, New York: John Wiley and Sons. 1974.

132

Farrer, Austin. Faith and Speculation. New York, New York: University Press, 1967.

Galilei, Galileo. “Letter to the Grand Duchess Christina of Tuscany, 1615.” Modern History Sourcebook, Last Modified Jan. 21, 2020, https://sourcebooks.fordham.edu/mod/galileo-tuscany.asp.

Galilei, Galileo. Two New Sciences. trans. Crew, Henry, de Salvio, Alfonso, Macmillan, 1914, https://www.stmarys- ca.edu/sites/default/files/attachments/files/Dialogues_Concerning_Two_New_Sciences.p df.

Goodenough, Ursula. “Who is a Religious Naturalist?” Theology and Science 15, no. 3 (August 2017): 231-234.

Graefe, Eva-Maria. “Quantum Chaos on Display.” Physics 6, no. 9 (January 22, 2013), http://link.aps.org/doi/10.1103/Physics.6.9.

Heisenberg, Werner. Physics and Philosophy: The Revolution in Modern Science. Amherst, New York: Prometheus Books, 1999.

Heisenberg, Werner. The Physical Principles of the Quantum Theory. New York, New York: Dover Publications, 1949.

Heisenberg, Werner. “Science and Religion”. Accessed 11/23/2020 https://www.edge.org/conversation/science-and-religion.

Jeans, Sir James. Physics and Philosophy. New York, New York: Dover Publications, 1981.

McGrath, Alister E. Science & Religion: An Introduction. Oxford, UK: Blackwell Publishers, 1999.

Moltmann, Jürgen. God in Creation: An Ecological Doctrine of Creation. London, SCM Press Ltd., 1985.

Moltmann, Jürgen. “God is Unselfish Love.” In The Emptying God: A Buddhist-Jewish- Christian Conversation, Edited by John B. Cobb Jr and Christopher Ives, 116-124. Eugene, Oregon: Wipf & Stock Publishers, 1990. https://books.google.ca/books?id=L39LAwAAQBAJ&lpg=PP1&dq=The%20Emptying %20God%3A%20A%20Buddhist-Jewish- %20Christian%20Conversation%2C&pg=PA116#v=onepage&q&f=false.

Moritz, Joshua M. Science and Religion: Beyond warfare and Toward Understanding. Winona, Minnesota: Anselm Academic, 2016.

133

Nancey Murphy. “Divine Action in the Natural Order: Buridan’s Ass and Schrödinger’s Cat.” In Chaos and Complexity: Scientific Perspectives on Divine Action, Edited by Robert John Russell, Nancey Murphy, and Arthur R. Peacocke, 325-357. Vatican City State: Vatican Observatory, 1995; Berkeley, California: Center for Theology and the Natural Sciences, 1995.

Newton, Isaac. The Principia: Mathematical Principles of Natural Philosophy. Translated by I. Bernard Cohen and Anne Whitman. Berkely, Los Angeles, London: University of California Press, 1999.

Peacocke, Arthur. Theology for a Scientific Age: Being and Becoming – Natural, Divine, and Human. Minneapolis, Minnesota: Fortress Press, 1993.

Pearcy, Nancy R., Charles B. Thaxton. The Soul of Science: Christian Faith and Natural Philosophy. Wheaton, Illinois: Crossway Books, 1994.

Planck, Max. Scientific Autobiography and Other Papers. Translated by Frank Gaynor. London: Williams & Norgate LTD., 1950.

Polkinghorne, John. Belief in God in an Age of Science. New Haven, London: Yale University Press, 1998.

___. Beyond Science the Wider Human Context. Great Britain: Cambridge University Press, 1996.

___. Exploring Reality. New Haven, London: Yale University Press, 2005.

___. Faith of a Physicist. Minneapolis Minnesota: Fortress Press, 1994.

___. Faith Science & Understanding. New Haven, Connecticut: Yale University Press, 2000.

___. God’s Action in the World. CTNS Bulletin, 10, no. 2, Spring, 1990.

___. “The Laws of Nature and the Laws of Physics.” in Quantum Cosmology and the Laws of Nature: Scientific Perspectives on Divine Action, Edited by Robert John Russell, Nancey Murphy, and C. J. Isham, 429-440. Vatican City State: Vatican Observatory, 1999; Berkeley, California: Center for Theology and the Natural Sciences, 1999.

___. “The Metaphysics of Divine Action.” In Chaos and Complexity: Scientific Perspectives on Divine Action, Edited by Robert John Russell, Nancey Murphy, and Arthur R. Peacocke, 147-156. Vatican City State: Vatican Observatory, 1995; Berkeley, California: Center for Theology and the Natural Sciences, 1995.

___. One World. Princeton, New Jersey: Princeton University Press, 1986.

134

___. “Physical Process, Quantum Events, and Divine Agency.” In Quantum Mechanics, Edited by Robert John Russell, Philip Clayton, Kirk Wegter-McNelly, and John Polkinghorne, 181-190. Vatican City State: Vatican Observatory, 2001; Berkeley, California: Center for Theology and the Natural Sciences, 2001.

___. Quantum Physics and Theology: An Unexpected Kinship.” New Haven Connecticut: Yale University Press, 2007.

___. The Quantum World. Princeton, New Jersey: Princeton University Press, 1984.

___. “The Quantum World.” In Physics, Philosophy and Theology, Edited by Robert John Russell, William R. Stoeger, S.J. and George V. Coyne, 333-342. Vatican City State: Vatican Observatory, 2000; Berkeley, California: Center for Theology and the Natural Sciences, 2000.

___. Quarks, Chaos, & Christianity. New York, New York: Crossroad Publishing Company, 1994.

___. Reason and Reality. Valley Forge, Pennsylvania: Trinity Press International, 1973.

___. “Response to Wesley Wildman’s ‘The Divine Action Project’.” Theology and Science 2, no.2 (2004): 190-192.

___. Science and Creation. Boston, Massachusetts: Shambhala Publications, 1988.

___. Science and Providence: God’s Interaction with the World. Shambhala Publications, Boston, 1989.

Rhodes, Richard. The Making of the Atomic Bomb. New York, London: Simon & Schuster, 1986.

Rohr, Dave. “Beyond Dead Silence: The Need for Cataphatic Affirmation Within Wesley Wildman's Apophatic Religious Naturalism.” Religious Naturalism Resources. Accessed June 2, 2020. http://people.bu.edu/wwildman/relnat/ Note that Wesley Wildman is the editor of this website.

Russell, Colin. “Science and Faith in the Life of Michael Faraday.” The Faraday Institute for Science and Religion. Last modified April 2007. http://faraday.stedmunds.cam.ac.uk/resources/Faraday%20Papers/Faraday%20Paper%20 13%20Russell_EN.pdf.

Saunders, Nicholas. Divine Action & Modern Science. Great Britain: Cambridge University Press, 1996.

135

Saunders, Nicholas. “Does God Cheat at Dice? Divine Action and Quantum Possibilities”. Zygon 35, no. 3 (September 2000): 517-544.

Schilpp, Paul Arthur. Albert Einstein, Philosopher – Scientist. LaSalle, Illinois: Open Court, 1970.

Serway, Raymond A. Physics for Scientists and Engineers. Philadelphia, Fort Worth, Chicago, San Francisco, Montreal, Toronto, London, Sydney, Tokyo: Saunders College Publishing, 2000.

Smedes, Taede A. “A Critical Evaluation of the Use of Chaos Theory in Theology.” In Studies in Science and Theology 8. Yearbook of the European Society for the Study of Science and Theology 2001-2001, Edited by N. H. Gregersen, U. Görman, H. Meisinger. Aarhus, 277-294. Denmark: University of Aarhus, 2002.

Smedes, Taede A. “Is Our Universe Deterministic? Some Philosophical and Theological Reflections on an Elusive Topic.” Zygon 38, no. 4 (December 2003): 955-979.

Spinoza, Benedict de. A Spinoza Reader: The Ethics and Other Works. Edited and translated by Edwin Curley. Princeton, New Jersey: Princeton University Press, 1994.

Stoeger, William R., S. J. “Conceiving Divine Action in a Dynamic Universe.” In Scientific Perspectives on Divine Action: Twenty Years of Challenge and Progress, Edited by Robert John Russell, Nancey Murphy, and William R. Stoeger, S.J., 225-254. Notre Dame, In.: University of Notre Dame Press, 2008.

Stoeger, William. “Describing God’s Action in the World in Light of Scientific Knowledge of Reality.” In Chaos and Complexity: Scientific Perspectives on Divine Action, Edited by Robert John Russell, Nancey Murphy, and Arthur R. Peacocke, 239- 261. Vatican City State: Vatican Observatory, 1995; Berkeley, California: Center for Theology and the Natural Sciences, 1995.

Stone, Jerome A. “Religious Naturalism and the Religion-Science Dialog: A Minimalist View.” Zygon 37, no. 2 (June 2002): 381-394.

Templeton Prize, 2020. John C. Polkinghorne – 2002 Press Release. [online] Available at: https://www.templetonprize.org/laureate-sub/polkinghorne-press-release/ [Accessed 24 November 2020].

Tipler, Paul A. Physics for Scientists and Engineers. New York, New York: Worth Publishers, 1991.

Wigner, Eugene. Symmetries and Reflections: Scientific Essays Cambridge: MIT Press, 1970.

136

Wigner, Eugene. “The Unreasonable Effectiveness of Mathematics in the Natural Sciences.” In Communications in Pure and Applied Mathematics 13, no. I (February 1960).

Wildman, Wesley J. “A Counter Response on ‘The Divine Action Project’.” Theology and Science 3, no.1 (2005): 17-29.

Wildman, Wesley J. “The Divine Action Project, 1988-2003.” Theology and Science 2, no.1 (2004): 31-75. file:///E:/JFB/Theology/Ph.%20D.%20Studies%20at%20U.D/Dissertation/Wildman/Wild man%20DAP.pdf.

Wildman, Wesley J. “Further Reflections on ‘The Divine Action Project’.” Theology and Science 3, no.1, (2005): 71-83.

Wildman, Wesley J. “Ground-of-Being-Theologies.” In The Oxford Handbook of Religion and Science, Edited by Philip Clayton, Oxford, England: , 2020. https://www.oxfordhandbooks.com/view/10.1093/oxfordhb/9780199543656.001.0001/ox fordhb-9780199543656-e-37.

Wildman, Wesley J. “Incongruous Goodness, Perilous Beauty, Disconcerting Truth: Ultimate reality and Suffering in Nature.” Accessed June 2, 2020, https://www.google.com/search?q=INCONGRUOUS+GOODNESS%2C+PERILOUS+B EAUTY%2C+DISCONCERTING+TRUTH%3A+ULTIMATE+REALITY+AND+SUF FERING+IN+NATURE+Wesley+J.+Wildman&oq=INCONGRUOUS+GOODNESS%2 C+PERILOUS+BEAUTY%2C+DISCONCERTING+TRUTH%3A+ULTIMATE+REA LITY+AND+SUFFERING+IN+NATURE+Wesley+J.+Wildman&aqs=chrome.0.69i59. 1683j0j7&sourceid=chrome&ie=UTF-8.

Sources with no author given:

“Ian G. Barbour, Winifred and Atherton Bean Professor Emeritus of Science, Technology and Society, Carleton College.” The Gifford Lectures: Over 100 Years of Lectures on Natural Theology. Accessed on 7/3/2018, https://www.giffordlectures.org/lecturers/ian-g- barbour.

137