AnthonyFlood.com

Panentheism.  Revisionism.  Anarchocapitalism.

 

David Ray Griffin

[link to CV]

Home

Essays by Me

Essays by Others

From Physics and the Ultimate Significance of Time: Bohm, Prigogine, and Process Philosophy, edited by David Ray Griffin, Albany: The State University of New York Press, 1986, Chapter 1, “Introduction,” 1-48

Time and the Fallacy of Misplaced Concreteness

David Ray Griffin

1.1. Time, Physics, and Metaphysics

In the title of this book, and especially in this introductory essay, “time” means time as known and understood in human experience.  What exactly this is, of course, is by no means obvious!  But there is considerable consensus, even among writers who disagree radically about the ultimate significance of time so understood, that time as experienced involves at least the following characteristics: (1) a one-way direction that is in principle irreversible, (2) categorical differences between past, present, and future, and (3) constant becoming.

It is commonly held that physics provides no basis for any of these features.  The so-called fundamental laws of physics are often said to require “time” only in the very abstract sense of the t-coordinate, on which events are strung out.  Not only do they provide no basis for a distinction between past, present, and future, and a sense of an irrevocable ongoing; they do not (with possibly one minor exception)1 even allow for “anisotropy,”2 the phenomenon in which the order of events when read off in one direction can be objectively distinguished from the order when read off in the other direction.  And it is often considered entirely a matter of convention that the vector points from the “past” to the “future” (Davies, 1976, 12f).

Thermodynamics does provide for anisotropy and hence, as K. G. Denbigh (1981, 167) says, brings “the concept of ‘time’ a little closer to what we have from our own experience.”  The thermodynamic “gradient of monotonically changing entropy states . . . appears to correlate completely with our own judgement of the temporal order of events.”  However, according to the dominant interpretation (at least prior to Prigogine’s influence), there is still no distinction of past, present, and future and no “one-way” going.  “For although thermodynamics finds the two directions of time to be distinguishable, it does not display the one direction as being in any sense ‘more real’ than the reverse direction. . . . The question ‘Which direction along the t-coordinate is the real direction?’ just doesn’t arise in physical science” (Denbigh 1981, 167).  Furthermore, the anisotropy provided by thermodynamics is said to be a property of large numbers of entities or events, not of these individual entities or events themselves, and hence it is considered a statistical matter.3  Even at this level there is said to be no irreversibility in principle.4  

In time as experienced, however, that today comes after yesterday, that earlier events come before later ones and not after them, is not a matter of interpretation of statistical data.  Rather, the irreversibility of time seems to be analytic, a necessary aspect of what we mean by time.5  There seems to something qualitatively different about time as we experience it and “time” as defined by theoretical physics or even thermodynamics,6 according to heretofore dominant interpretations. 

In time as we experience it, the past, present, and future are categorically distinct.  The past is composed of all those events which are totally determinate, in which all options have been eliminated.  As the epigraph says, once time’s moving finger has writ, neither piety nor wit can cancel a word of it.  The future, on the other hand, is that realm in which at least some indetermination remains; decisions remain to be made from among various possibilities.  Although absolute determinists say that there are in reality no alternative possibilities, beyond what will actually occur, everyone in practice lives as if the future is in some sense open, as if there are real decisions to be made—which brings us to the present.  The present is the realm in which decisions are being made: some possibilities are being turned into actualities while other possibilities are being excluded from actualization.  So, we have these categorical differences: the past is the fully determined, the present is what is becoming determined, and the future is the still-to-be-determined.  In classical physics there was no basis for these categorical distinctions.  Some now say that quantum physics provides a basis for this tripartite division,7 but this issue is still debated.  The dominant view has been that fundamental physics is timeless. In any case, few thinkers deny that human experience of time includes this threefold division into past, present, and future—or at least the illusion of it.8

The term constant becoming is used to point to the fact that the present, or “now,” does not “stand still,” as some put it.  More precisely, the present “now” never divides the same sets of events into past and future.  In each new “now” there are events in the past that were not there before and that previously had been at most anticipated as possible, or perhaps probable, events.  This is the feature of time that has been asserted by many writers to be most totally absent from physics.  Physical theory is usually said to be indifferent to any idea of becoming, of events previously “in the future” coming into present existence.9  This would be all the more the case if this idea were taken to include the idea that this coming into existence involves the transformation of potentialities into actualities.

These three widely agreed-upon aspects of time as we experience it,10 at least as I have explicated them, all presuppose each other.  Hence, none of them really provides any “stronger” sense of time than the others.  For example, part of the reason that time is irreversible is that the past is composed entirely of totally settled, determinate events; it would be self-contradictory for those events somehow to come to be in the future, for that would mean they would be in the realm of the still partially indeterminate.  Likewise, it is because the present involves the transformation of possibilities into actualities that there is constant becoming, which means that time cannot “stand still.”  Once some possibilities have been chosen and others excluded, one cannot face these same options.  Those decisions are, as we say, now “behind us,” and we now confront new options.

This book is concerned with the ultimate significance of time in this sense.  The epigraphs at the head of this chapter were chosen to illustrate several points.  One is that humanity has harbored greatly different attitudes about time, with its irreversibility and incessant becoming.  Some people focus on the aspect of time that leads them to call it a “devouring tyrant”; others see it as a “refreshing river.”  Some see it in theodical terms, as ultimately healing all wounds and wounding all heels; others see its “perpetual perishing” as constituting the ultimate problem of evil.  Some passionately believe the universe would be an aesthetic failure unless it were symmetrical in all respects; others with equal passion believe that unless the relations of the present to the past and to the future are asymmetrical, so that we have genuine freedom partially to create the present and hence the future, life would be meaningless.  Some believe that a life of wisdom and happiness requires coming to terms with the ultimate significance of time; others see it as requiring an appreciation of time’s insignificance, even illusoriness.  These various attitudes toward time are not directly the subject of this volume.  However, they are not irrelevant to the subject, since what we want to believe, for moral, religious, or aesthetic reasons, usually conditions the types of evidence to which we attend and the relative weight we give to the various considerations.

A second purpose of the epigraphs is to illustrate the point that thinkers differ greatly on the subject of this volume: the ultimate significance of time.  Finally, many of the epigraphs directly reflect the more complete topic of this volume:  the bearing of physics upon the question of the ultimate reality of time.  Here again, thinkers differ drastically. They can be divided into three major groups.

First, there are those who hold that time is to be ascribed no other nature, and no more reality, than it has in the fundamental laws of physics.  This position has had much strength in the last two centuries, because of widespread acceptance of the metaphysical position of reductionistic materialism among physicists and philosophers (even if it was not acknowledged as a “metaphysical” position) and because of widespread abdication on the part of philosophers from involvement in the philosophical issues in physical theory.  The majority of philosophers who have concerned themselves with these issues, the “philosophers of science,” have done so from a viewpoint (e.g., operationalism) that has sanctioned the tendency of physicists to give terms no larger meaning than that vouchsafed by the methods, data, and theoretical formulations of their own discipline.  In any case, those holding this position have tended to speak of time as an illusion (see the epigraphs by Russell, Einstein, de Broglie, Gold, de Beauregard, Griinbaum, Weyl, Mehlberg, and Davies).11  

A second position is constituted by those who agree with those in the first group that time, like all meaningful concepts, must be defined by science, but who hold that, although the concept of time is not grounded in the fundamental laws of physics, it is grounded in some other feature of the world that is susceptible to scientific treatment.  The most popular form of this position is the view of time as rooted in the laws of thermodynamics.  From this point of view, it is arbitrary to say that time is unreal simply because it is not reflected in the “fundamental (timeless) laws” of the universe.  Science also deals with historical contingencies and so can include the “initial conditions” of the universe (which are presumed to be the source of “time’s arrow”) among its fundamental data.  From this viewpoint, it is fundamentally the direction of entropy that defines the reality of time.  Our own experience of time is taken to be derivative from the fact that we are composed through and through of entities involved in the universal entropic process.12

However, other approaches are possible within this general position.   Time could be regarded as first arising in living processes and hence would be established by the biological sciences.  Or it could be thought a mind-dependent property and hence to have arisen only when organisms with minds sufficiently analogous to our own had emerged (often thought to occur only with the development of a central nervous system).  Proponents of these views would tend to stress the reality of time to the degree that they consider living, or psychological, processes as genuinely real—not some frothy epiphenomena on the surface of the really real, strictly physical, processes.

In fact, within this second position as a whole the sense in which time is spoken of as real is a matter of degree, seeming to depend not only upon the author’s systematic position but also upon his or her desire to regard the universe as essentially temporal or timeless.  None of those in this group regard time as of strictly ontological or metaphysical significance, i.e., as rooted in the very nature of reality.  They all see it as something contingent, as something that depends upon particular features of our universe that conceivably could have failed to occur—or at least they believe that they as scientists can only so regard time.  Since time is therefore regarded as a contingent feature of reality, it can be regarded as more or less “real” or “illusory,” depending upon the other interests and biases of the author.  Hence, the difference between this and the first position is ultimately a difference of degree, so the question of which writers would be classed in the second as opposed to the first position is somewhat arbitrary.  However, it is a difference of degree that becomes virtually a difference in kind, as some advocates of this second position stress the reality, universality, and significance of time so strongly as to make their difference from the third position discernible only by careful reading.  This would be true, for example, of Ilya Prigogine—if he is to be understood as exemplifying the second rather than the third position (see section 1.3).

The third position is constituted by those who hold that it is not the task of one of the special sciences to define and account for the reality of time.  For one thing, the methods adopted by the sciences are, at least thus far, said to be such as to preclude treatment of the temporal features of reality; hence, looking to physics to learn about time is like asking a computer to answer a question it has been programmed to ignore.  Furthermore, every attempt to explain temporality in terms of some physical characteristic of the world involves, it is claimed, the fallacious procedure of reducing a more fundamental concept to a less fundamental one (see epigraphs by Gale, Bridgman, and Whitrow).  Within this third group, some would say that time is simply a “primitive” concept that need not and probably cannot be explained in any sense, either physically or metaphysically.  It suffices to see that time (as, e.g., Gale would claim) is an inevitable feature of our experience, as reflected in our ordinary language, which is presupposed by all our technical languages, so it is senseless to try to explain it away on the basis of any of those technical languages.  However, some within this third group would say that time is a topic for metaphysics, or ontology,13 meaning thereby that an approach is required that does not limit itself to the methods and abstractions of any of the special sciences, but that attempts to synthesize the presuppositions and results of all the special sciences with each other and with the knowledge and especially presuppositions of human experience in its fullness (which may include features not in the province of any of the special sciences).

 

1.2. Time in Process Philosophy

There are many ontologies or metaphysical systems that fit the general characterization given in the previous sentence.  (There are also some that do not, since they ignore the results of the special sciences; but they are themselves ignored here, just for that reason.)  This introductory essay is written from the point of view of one of them, the “process philosophy” derived primarily from the writings of Alfred North Whitehead.  From this perspective, the natural sciences, at least as practiced thus far, methodologically abstract from the full concreteness of the entities or processes they study.  Therefore, to jump from the mere fact that time is not present in natural science to the conclusion that time is not real at the fundamental level of nature is to commit the “fallacy of misplaced concreteness.”  The fallacy is to treat the abstractions from certain things—as abstractions focused on because of certain interests and methods—as if they were the concrete things themselves.14  It is to treat the map as if it were the territory, assuming that what is not on the map is not in the actual terrain itself.

The thesis implied by the title of Whitehead’s major book, Process and Reality, is that process is the concrete reality of things and, conversely, that concrete realities are processes.15  From this viewpoint, time, or temporality, is an ultimate feature of reality.  It is not itself an actual or concrete entity; it is a relation—a relation of conformity to and inclusion of the past.  This is a version of the relational theory of time.  However, many relational theories make time a contingent feature of reality, since time is presumed to depend upon a kind of relation that may or may not exist.  For example, time in Augustinian theology—according to which the world was created ex nihilo in a sense that made the existence of finite things as such purely contingent, dependent upon the voluntary decision of God—was purely derivative, a relation among such contingent things.  Of course, if the everlastingly existent creator had itself been a temporal being, as in Isaac Newton’s thought (as Milič Čapek points out),16 then time could be considered to be a relation and yet everlastingly real, as it would most fundamentally be the relation between the creator’s successive states of consciousness.  (Most commentators have ignored Newton’s heterodox theology, and his talk of “absolute time” has generally been misunderstood to mean that time is not in any sense a relation and hence can exist apart from actual events.)  But since traditional theology held God to be nontemporal in every sense, time was itself considered part of the contingent creation.  Much modern science-based philosophizing, in thinking of time as a derivative, emergent reality, has thereby held to this fundamental dictum of traditional theology, that time or process has no ultimate reality.

But process philosophy, while it speaks of the present world as creation, emphatically rejects this notion of creatio ex nihilo, according to which there might be no plurality of things whatsoever, and the related idea that temporality or process is a derivative feature of the world.  Although time is a relation, there have always been processes or events having the twofold structure of inclusion and anticipation (i.e., anticipation of being included by future events). The existence of such events or processes is hypothesized to be ontologically necessary, simply part of the ultimate nature of things, which just is.17  In answer to the question, for example, What existed before the “big bang?” process philosophy would claim—assuming the correctness of this theory of the origin of our universe—that such events or processes existed.   Hence, process philosophy can be characterized as pantemporalism.

From this perspective, the relations and objections of process philosophy to the other basic options for the ultimate reality of time can be clarified.  The three options would be: reductionistic nontemporalism, dualism, and pantemporalism.  Reductionistic nontemporalism can be summarized in the following three propositions.  (1) The fundamental laws of physics tell us all the fundamental characteristics of the universe.  (2) These laws do not include time.  (3) Therefore time is unreal, an illusion fabricated and projected onto reality by human experience.18  

One objection of process thought to this reductionistic position has already been suggested: there is no good reason to suppose that the fundamental laws of physics, especially of physics in its infancy (a few hundred years is not very long), give us all the fundamental characteristics of the universe, even of the most “elementary”19 entities or processes constituting it.  There is much more reason to suppose that physics, at least as developed thus far, involves significant abstractions from the concrete realities of things.  A second objection is that, according to the reductionistic viewpoint, human experience, with its inherently temporal structure, is an epiphenomenon, not a full-fledged actuality; and yet it is credited with the power of creating the illusion of time.  This attribution of such enormous creative power to unreal things should be taken as a reductio ad absurdum of reductionism.  A third objection is based on an idea that process philosophy shares with the “commonsense” and “pragmatist” schools of philosophy.  This is the idea that the ultimate criteria for testing philosophical doctrines are those notions that all people in fact presuppose in practice, even if they deny them verbally.  I call these the “hard-core commonsense” notions, to distinguish them from those ideas which are often thought of as “common sense” but which belong to the “soft core” of one’s repertoire of common sense (that is, they are really rather provincial ideas and are definitely not ones that all human beings inevitably presuppose in all they do).  Because the hard-core commonsense notions cannot be consistently and hence meaningfully denied—since the very attempt to refute them would presuppose them—they should provide the ultimate constraint upon philosophical opinions.20  We should accept no doctrine in our theory that we cannot consistently live by in practice (as the pragmatists put it).

The doctrine of the ultimate reality of time is one of those doctrines.21  The very act of living, including the activity of articulating and defending scientific and philosophical ideas, necessarily presupposes the dimensions of time as described above: that we can affect the future, but not the past; that there is an irreversible directionality to reality; that the present involves turning some possibilities into actualities and excluding others; and that—however inadequately this formulation puts it—the present “now” constantly moves.  Rejecting this temporal aspect of reality as an illusion is done on the basis of accepting the truth of some other proposition as the major premise for a syllogistic argument.  For example, the proposition used above was:  “The fundamental laws of physics tell us all the fundamental characteristics of the universe.”  This is obviously not one of the “common notions” that all human beings inevitably presuppose in practice.  On what grounds do we say that we are so certain that this proposition is true that we will deduce from it the falsity of a proposition that is among the universal presuppositions of human practice?22  

There is a fourth basis on which process philosophy considers reductionistic nontemporalism unworthy of belief: the widespread acceptance of a materialistic metaphysics and of a corresponding sensationalistic epistemology, according to which anything knowable is knowable through the physical senses, has led to the supposition that all meaningful concepts apply only to objects of sensory experience.  But our ordinary notion of time is emphatically not derived from sensory experience; we do not see (hear, smell, taste, or touch) the past, we do not see unactualized possibilities, we do not see a moving “now.”  Since from a sensationalist perspective, as Richard Gale points out (1968, 69f.), the factual content of all assertions can be equated with their sensible content, it is assumed that assertions about time can be stripped of those features arising from nonsensory experience without loss of anything essential.  However, Whitehead (1966, 72-75, 133, 152-62) argues that scientific theory presupposes a set of concepts, time being only one among them (“causation” is another), that is rooted in a kind of perception (which he calls “prehension,” or “perception in the mode of causal efficacy”) that is not sensory and that is in fact more basic than sensory perception.  Sensory perception provides precision of observation, but not the meaning of the concepts employed in scientific theory.  Once it is realized that scientific theory in general cannot survive on the basis of a rigidly applied sensationalist epistemology, there should be less compulsion to regard as “phoney”23 those characteristics of time that are derived from nonsensory aspects of experience, such as memory and anticipation.

The second position in contrast to pantemporalism, dualism, might at first glance seem the most reasonable.  According to this view, there is no time, no temporal process, at the level of the most elementary entities constituting nature, such as that of individual atoms and subatomic particles.  Time is said to emerge only as a result of the emergence of certain conditions or types of events.  Some dualists, as indicated earlier, locate this at fairly high levels, e.g., with the emergence of life, consciousness, or, in extreme cases, human consciousness.  But others locate the emergence of time at a much more primitive level, in some cases so primitive that to insist upon a distinction between this view and that of pantemporalism may seem like quibbling.  However, there is an important philosophic problem involved—the same problem that plagues every form of ontological dualism.  This is the problem of how two ontologically different types of entities can interact.   The best-known form of ontological dualism is the one responsible for the “mind-body” problem, the Cartesian dualism between experiencing mind and non-experiencing matter.  Most philosophers have given up this view of the mind-body relation precisely because it has proved impossible to understand how mind and body, so understood, could interact (at least without resorting to the appeal, less acceptable now than in the seventeenth and eighteenth centuries, to a divine omnipotence that is not deterred by mere impossibilities).  But the problem of understanding how temporal and nontemporal things could causally interact (emergence presupposes a type of causal influence) is at least as difficult, as Čapek points out.24  How can we even think coherently about events for which the distinction between past, present, and future obtains as interacting with events (or things, since the term events already presupposes time) for which this distinction does not apply?  This difficulty is one of the major reasons process philosophers urge pantemporalism as a more reasonable view.25  

Process philosophy’s pantemporalism is articulated in several essays in this volume, including those by Bjelland, Hurley, Stapp, Cobb, Čapek, and me.  However, the crucial point, mentioned earlier, needs to be developed here.  For Whitehead, the reality of time, with its irreversibility, is based on the fact that the actual world is composed exhaustively of momentary events that include, partially but really, preceding events, which had in turn included previous events, and so on back.  In Whitehead’s words: “This passage of the cause into the effect is the cumulative character of time.  The irreversibility of time depends on this character” (1978, 237).

Because of this character of events, a present event is not independent of previous events; rather, it presupposes just those events, since it includes them and is largely constituted by this inclusion.  Because of this essential inclusion of prior events, the idea of successive events occurring independently of each other arises only by abstracting from the full reality of the events.

Time in the concrete is the conformation of state to state, the later to the earlier; . . . pure succession is an abstraction of the second order, a generic abstraction omitting the temporal character of time. . . . The immediate present has to conform to what the past is for it, and the mere lapse of time is an abstraction from the more concrete relatedness of “conformation.”  (Whitehead, 1959, 36)

It is therefore a fallacy to think of the real events or things making up the world as having the property of “simple location,” which would mean that they could be satisfactorily described without reference to prior and following events.26  This is no small point, since the traditions of Humean and thereby Kantian philosophy have presupposed just this idea of events as “simply occurring.”27  That is, the events in themselves were held to have no inherent conformal relationship to prior events.  These philosophical traditions are thereby based on the fallacy of misplaced concreteness.

The same is true of those materialisms or dualisms that think of nature as composed of “self-contained particle[s] of matter.”28  Rather than thinking of enduring particles as the fundamental entities of the world, Whitehead sees each enduring object as composed of a rapidly occurring series of events, each of which includes aspects from its predecessors in that enduring object but also aspects from other prior events as well.  Hence, Whitehead opposes the widespread view that an individual atom is timeless; rather, it is a “temporally-ordered society” of actual occasions.29  Again, the mistake is to confuse an abstraction (the form of energy, which remains relatively the same through time) with the concrete reality, which is a series of momentary events, each of which derives its form largely through conformity with predecessors and passes it on to successors.  Since an individual atom (or even electron) has a temporal structure, time or temporality does not first arise as a statistical effect of the interactions among a multiplicity of atoms.

I previously drew an analogy between three views on time and three views on the mind-body problem.  I now suggest that the relation between them is more than an analogy.  The two issues are finally one.  I will introduce this idea by reference to the thought of a philosopher who argues for the ultimate unreality of time in the physical world.  Adolf Grünbaum contends tirelessly that time, in the sense of becoming, is a mind-dependent property.  From this he concludes that it cannot exist in the physical universe.  However, although he sometimes speaks as if he meant only human minds, since he speaks of time in the sense of becoming as “anthropocentric,” in more careful formulations he makes clear that he does not limit the requisite mind to that of human beings (Grünbaum, 1967, 152, 179-80).  Although most process thinkers see Grünbaum as one of their staunchest opponents,30 as indeed he is in most respects, I want to stress the agreement between his thought and process thought on the mind-dependent nature of time.  Of course, process philosophers generalize “mind” much more broadly than Grünbaum would countenance; he is hesitant even about cockroaches (Grünbaum, 1967, 179f.), whereas process philosophers generalize “mindlike”—or, better, “experience”—to all actualities whatsoever.  We agree with Grünbaum that time is experience-dependent.  But since we hold that all actualities are units of experience and that a plurality of such actualities necessarily exists, we hold that time exists necessarily, not as a contingent emergent.  (“Actualities” here refers to genuine individuals, not aggregates; see 9.3.2, below.)

Accordingly, process philosophy’s pantemporal-ism and its panexperientialism are ultimately one and the same thing (as Andrew Bjelland’s essay makes clear).  Likewise, it is consistent for reductionistic materialism, which denies experience except as an epiphenomenal illusion, to deny time except as an illusion created by an illusion.  And it is consistent for dualists on the mind-matter issue to be dualists in regard to time.30a  Finally, process philosophy’s pantemporalist, panexperientialist position stands as a clear alternative both to reductionistic, materialistic nontemporalism, and to temporal-nontemporal, mind-matter dualism.  It holds that temporality as such and experience as such do not emerge at some point, that becoming is not somehow generated from nontemporal being, nor experience from nonexperiencing matter; indeed, it holds that such generations are unthinkable.  However, in contrast with reductionism, it does not thereby regard experience and time as illusions that magically appear at some point.  Accordingly, it denies that it is the task of a special science, whether physics or some other, to determine when and how time, or its illusion, first appeared.

The relevance of the fallacy of misplaced concreteness to this discussion of the relation between panexperientialism and pantemporalism must now be clarified.  Whitehead, in one of the quotations at the head of this essay, speaks of that science which finds no creativity (and hence no time) in nature as ignoring half the evidence.  He means this literally, for he regards each momentary event as having two modes of existence.  It comes into existence as a process of becoming (called “concrescence,” the process of becoming concrete).  In this mode of being it has experience—not usually self-consciousness, or even consciousness, but an emotional internalization of the environment nevertheless.  In this mode it is a subject.  This is what the event is in and for itself.  In this mode, it is known only to itself; it cannot, in principle, be perceived by any other subject.  However, this mode of existence is quickly over—it may last from (perhaps) a tenth of a second in some cases (e.g., in human experience) to a billionth of a second in others (e.g., in a subatomic particle that vibrates a half-billion times per second).31  When this process of concrescence is completed, the event begins its second mode of existence: it becomes an object for others, an object to be felt by subsequent events—which might include being perceived by a human observer or the instruments thereof.  In other words, it becomes a cause influencing later events.

Now, when Whitehead says that it is inherent in scientific methodology to ignore half the evidence, he means that such methodology considers things only as they are perceivable—either directly or by means of magnifying instruments.  This means that it considers things only when they are objects, after their period of subjectivity, or experience, is over.  In other words, it considers things only on their extrinsic side and ignores their intrinsic natures.  It may consider a “thing” as a process, but it deals only with its extrinsic process, which Whitehead calls “transition,” since there is a transition of energy from the one event to another.  But it ignores the event’s intrinsic process, which is “concrescence.”  In fact, as Whitehead states elsewhere, physics does not even deal with half the evidence, since it contains the extrinsic side of things only in regard to their spatiotemporal effects upon other things (1926, 220).  In any case, all the features of time discussed earlier are rooted in the intrinsic reality of events, in the process by which they become concrete, or determinate, for it is here that the event includes the past events into itself and it is this inclusion that makes time irreversible.32  Accordingly, any approach that commits the fallacy of misplaced concreteness by equating the extrinsic side of events with their complete reality will necessarily miss the roots of time in those events.

Behaviorism was the result of the decision that, to be a science, psychology had to ignore the intrinsic reality of human beings, which is known only by being one, in favor of looking exclusively at the features of human beings that can be known from without, through the sensory perception and magnifying instruments of the exterior observer.  Those who promoted this form of psychology tried to get along without using experience, consciousness, purpose, aim, desire, feeling, or any other “subjective” terms.  They wanted to exclude all “anthropomorphic” categories from the study of human beings, so that psychology could finally become a genuine science, like physics and chemistry.33  But now it is generally accepted that an exclusively behavioristic approach cannot deal with the most important features of human beings.  If we think of human experience as fully natural, why should it be excluded from psychology as a natural science?  The most adequate approach to the study of human beings comes from combining behavioral and introspectionist methods.

But if this point is accepted, and we are nondualists, should not the twofold approach be extended to all levels of actuality?  Of course, we cannot ask nonlinguistic entities what it feels like to be one of them in order to get at their intrinsic natures.  But likewise we need not study them on the basis of the assumption that they have no intrinsic side, that they have no experience, that they are nothing but objects.  For that assumption is pure speculation, no less “metaphysical” than its contrary.  By reasoning from analogy—the grounds of which have been greatly strengthened by acceptance of an evolutionary perspective—it is more reasonable to speculate that all individuals have an intrinsic nature, with some degree of experience, than to speculate that some of them do not, for we do; we know this directly.  I know that there is more to me than the behaviorist psychologist can describe.  And you know the same about yourself.  And you presume it about me in practice, no matter how solipsistically you may rhapsodize in theory.  Is it not less arbitrary to assume that this twofold mode of existence applies to all individuals, rather than to assume that it “emerged” at some point in the evolutionary process?  To assume the latter would be to assume an ontological dualism—no matter how vehemently one might reject the label.  To say that some things have an intrinsic as well as an extrinsic reality, and that other things have only an extrinsic reality—i.e., to say that some things are subjects as well as objects, whereas other things are only objects—is to affirm an ontological dualism, regardless of the name (such as “mind-brain identism”) with which one seeks to disguise it.34  

The widespread idea that physical events could in principle start running backwards presupposes the idea that temporal order is pure succession.  We know that it is nonsense to think of our own experience as running backwards because we know, at some level, that the order of our experiences is not that of pure succession, but of the partial derivation of later from earlier experience, with partial conformation of later to earlier. In Whitehead’s words:

Time is known to us as the succession of our acts of experience. . . . But this succession is not pure succession: it is the derivation of state from state, with the later state exhibiting conformity to the antecedent.  Time in the concrete is the conformation of state to state, the later to the earlier; and the pure succession is an abstraction from the irreversible relationship of settled past to derivative present.  (1959, 35)

Most people would agree that this correctly describes human experience, but they do not see how it could apply to “purely physical” processes.  Yet Whitehead suggests that it does: the deleted words in the above quotation are: “and thence derivately as the succession of events objectively perceived in those acts.”  Because he, on the basis of an evolutionary, nondualistic outlook, takes unitary “physical events” to be not different in kind from the “mental events” constituting our own immediately known experience, he can hold that time as “known to us in the succession of our acts of experience” can be attributed by analogy to “the succession of events objectively perceived in those acts.”  I put “physical events” and “mental events” in scare quotes to indicate that this dualistic language is mistaken from the perspective of process philosophy.  The difference between the proton and the psyche is one of degree, not of kind (in an ontological sense).  One who holds otherwise is a dualist, no matter how odious such a designation may be.

If this outlook were to become the accepted framework (or paradigm) within which science is practiced, in place of the ontological materialism and dualism between which most scientific work in the modern period has oscillated, one of two consequences would follow for the question of the reality of time in physics.  One possibility would be that physicists would retain the methodological limitation to the extrinsic nature of events.  Only now they would see this to be a self-imposed limitation and would assume that, apart from the aspect of events upon which they focus qua physicists, the events also have an intrinsic side, which is analogous in some remote way to our own experience and in which the basis for temporal relations at the most elementary level of nature is located.  Hence, they would be freed from the assumption that it is their task, qua physicists, to define the meaning of time and account for its existence—or, failing this, to deny its reality.

A second possibility would be that they would reconceive the nature of physics so that it could, in principle, take account of the intrinsic as well as the extrinsic nature of events.  If this occurred, the result could properly be called the beginning of “post-modern” science, since it can well be claimed that the most significant feature distinguishing modern from pre-modern science is the exclusion of all categories of subjectivity or experience from consideration, either as data or for use in explanatory theory.35  However, the resulting science would be postmodern, not a return to pre-modern, insofar as the gains in rigor and precision acquired during the modern period were retained.36  In such a science the reality of time, with its features as known through human experience, could be included within science itself.

 

1.3. Time in the Work of Bohm and Prigogine

I turn now to the question of how the programs of Bohm and Prigogine are related to the issues involved in physics and time as seen from the perspective of process philosophy.  Bohm seems closer to developing a post-modern physics in the above sense.  He, like Whitehead, understands apparently enduring particles to be “world tubes” composed of rapidly occurring series of events.  Furthermore, he sees each momentary event as enfolding or “implicating” the whole of reality within itself.  Finally, he says that physics thus far has stressed the “explicate order,” but that it now needs to deal also with the “implicate order.”  At least in some of his formulations, this implicate order seems to involve that phase of the momentary events in which they enfold the whole of reality within themselves.

Thus, Bohm would seem to be developing a physical theory corresponding to Whitehead’s metaphysical hypothesis that each momentary event has a subjective side and that this subjective side is a process of concrescence in which the whole past world, under abstraction, is included.  However, sometimes Bohm has spoken as if the “whole” that is enfolded in each event were not simply and directly the past events, but a nontemporal order in which the future as well as the past is contained.  If this were his position, the reality of time would be denied, for from the ultimate point of view (i.e., of the “superimplicate order”), what is future to us would already (eternally) be as determinate as what is past.  I will not discuss this and other issues in Bohm’s thought any further here, however, since they are explored in my essay and Cobb’s, below.

Prigogine’s position presents the process philosopher with something of the inverse situation.  On the one hand Prigogine rejects, at least so far, any revision of the basic nature of natural science so as to allow it to include subjectivity among its data or to speak of any aspect of events that is not in principle observable.  On the other hand he is as intent as any process philosopher on insisting upon the ultimate significance of time, and this has led him to develop a “post-classical” science.

Prigogine indicates that the problem of time has been at the center of the research he has been pursuing all his life (Prigogine and Stengers, 1984, 10).37  He says that the problem of the “two cultures” has not been that scientists have not read enough humanities and humanists enough science, but that there has been nothing in common between the two thought worlds.  At the root of the cleavage has been the fact that, while the humanities and social sciences are necessarily time-oriented, classical science has been nontemporal (xxvii).  Furthermore, this nontemporal view has been part and parcel of an alienating science that has portrayed a dead, debased nature, creating an inevitable opposition between humanity and nature (3, 4, 6, 89).  This portrayal of nature has led, by reaction, to antiscientific metaphysics (3, 7, 31 ff., 79).  Overcoming the dualism between physics and philosophy, between matter and life, between science and the humanities, between nature and humanity, requires an enlarged science, with a new idea of time (14, 96, 175, 298).  This in turn involves a new concept of matter, a return in a sense to one aspect of pre-modern naturalism, in which matter was seen as active and self-organizing (22, 32f., 36, 75, 82, 287, 291).  Matter now should be seen as capable of “perception” and “communication” (33, 145, 163, 171, 180, 181) and as manifesting behavior that depends upon its special past history (153, 161).  Furthermore, we should beware of speaking of “elementary” particles, since this term suggests an autonomy from context that is untrue (287).

All of this adds up to a post-classical science, taking “classical” science to be that which emphasized time-independent laws and deterministic processes, thereby portraying nature as a grand tautology (2, 77, 213).  The new science, which forges a new alliance between humanity and nature (to replace the animistic alliance broken by modern or classical science), stresses nondeterministic processes, in which there is intrinsic randomness—i.e., randomness due not only to our ignorance of deterministic causes (234, 239, 298)—for only genuinely nondeterministic processes can be irreversible (xxvii, 16, 276, 277).  The fundamental character of irreversible processes is Prigogine’s key thought; only this kind of process gives science the kind of time that is presupposed in all human experience, and hence in philosophy and the humanities in general.  Prigogine’s program is to reveal the existence of irreversibility at all levels, including and especially the microscopic level (16, 232, 258, 288, 289):  there must be something at the microscopic level that provides the root of irreversibility at the macroscopic level.  This latter irreversibility cannot intelligibly be thought to emerge, as a miracle, from fully reversible processes (285, 289, 298).

On all these points, Prigogine is in complete agreement with Whiteheadian process philosophy.  Indeed, he often stresses his agreement with Whitehead as well as with Bergson (93-96, 129), seeing his own task to be that of giving scientific content and precision to their metaphysical speculations (24, 310).  However, beyond these agreements there are some differences.  These differences can be read as merely methodological—reflecting the different criteria and resources to which Prigogine, qua physicist, and Whitehead, qua philosopher, can and must appeal—or the differences could be read as more radical.

I will begin with the more radical interpretation.  According to this reading, whereas Whitehead would exemplify the third position on the bearing of physics upon the ultimate reality of time (see above, section 1.1), Prigogine would be seen as exemplifying the second.  The crucial difference would be whether human experience as such were used to establish the fundamental meaning and nature of time.  For the later Whitehead (see Hurley’s essay, below) it is, and the irreversibility in nonhuman nature is accounted for by the postulate that it is composed of processes that are analogous to our experiential process.38  The adoption of this position by Whitehead was part of his move from the “philosophy of nature” in his pre-1925 works, in which the human subject is excluded and nature is regarded only as the object of sense experience, to “metaphysics,” in which experience and the experienced are to be integrated into one scheme of thought, and inquiry about nonhuman events is extended beyond the description and correlation of their appearances to the question of what they might be in themselves.

Prigogine, on the other hand, sometimes speaks as if a post-classical science can reconceive time and nature adequately without breaking with classical science’s dictum that “the soul which counts” lies outside the province of physical science (Prigogine and Stengers, 1984, 22).  He sometimes (e.g., near the end of the paper in this volume) seems to account for our experience of time on the basis of the fact that we are examples of those highly unstable dynamic systems in which randomness and hence irreversibility arise—which seems to make his procedure the opposite of Whitehead’s.  Consistent with this interpretation of Prigogine’s position is the fact that he sometimes appears to regard time as a contingent rather than universal feature of the physical world because it requires a minimum complexity (16, 239, 251, 298, 301).  There was, he says, a movement “from being to becoming”; i.e., becoming or time-irreversibility originated in the first stages of our universe (xxx, 278, 298, 300, 310).  Hence, the forward direction of things is only a tendency, not a necessary truth (xxvii, 300).  The world, accordingly, is pluralistic (xxvii, 251), or what I earlier labeled “dualistic,” involving a mixture of reversible and irreversible processes (xxvii, 251, 232, 258, 289).

Under this interpretation, Prigogine would differ only in degree from previous thinkers who regarded the task of physical science to be to establish the meaning of time and who considered lime a contingent feature that arose at some point in the development of our universe.  Prigogine would, of course, differ greatly in degree, since he stresses the priority of irreversible processes, in contrast to those who have regarded irreversible processes as exceptional and hence have had to see the origin of life, with its irreversible processes, as a virtual miracle.  Prigogine stresses that irreversibility occurs at every level of nature (285) and that reversibility is based on highly artificial, simplified situations: nature in the raw is not simple even at the most fundamental level, and in nature irreversibility is the rule, not the exception (8, 9, 10f., 215f.). 

But even on this point (to continue this interpretation of Prigogine’s meaning), the difference with Whitehead, while it may seem subtle, is important.  For Whitehead, reversibility is the result not of the artificial, but of the abstract.  That is, there is no true reversibility, not even in artificial, isolated conditions.  The idea that time could be reversed results from mistaking abstractions for the fully concrete entities.  In Prigogine’s framework, however, it seems that genuine reversibility can occur: in artificially constructed contexts there can be processes that are fully deterministic, hence symmetrical and reversible.  If time is defined in terms of the direction of physical processes, one would have to say that in those situations time itself is reversible.

The crucial question in regard to the correctness of the above interpretation of Prigogine’s position is just this, i.e., whether he means that the far-from-equilibrium processes provide the basis for a definition of time.  If he does, then, even though he says that the entropy barrier is infinite so that reversibility could never in fact occur, his position would suffer from the paradoxes afflicting all the other attempts to define time in terms of physical processes (see note 4, above).

However, there is another way to interpret Prigogine’s position.  This interpretation is based on his statement that the distinction between past and future is a primitive, i.e., prescientific, concept, which science must simply presuppose (1973, 590; 1980, 213).  Under this interpretation, Prigogine’s view would be not that it is science’s task to provide an adequate concept of time, but that science must become consistent with the primitive “commonsense” notion of time we have from our own experience.  Prigogine’s work would move physical theory toward this consistency by showing that reversible time is not a good tool even for complex dynamical systems and that irreversibility is primary at every level of reality.

In harmony with this interpretation, the movement “from being to becoming” would not be a physical one (which would make irreversibility contingent), but a conceptual one, signifying the move from a physical theory in which simple, reversible systems have an autonomous and fundamental status within complicated systems to a theory in which they become singular limiting cases of an asymmetrical model.  The pluralistic “mixture” of which Prigogine speaks would not refer to a true or ontological mixture, but would reflect his desire to make sense of the phenomenological difference that made classical dynamics possible, i.e., that some systems in fact behave—as a first approximation—in a reversible way.  Finally, under this interpretation Prigogine’s “artificial” would be the same as Whitehead’s “abstract.”  That is, “artificial” systems would be those which had been conceptually simplified to make them conform to the conceptual tools of a reversible dynamics.

According to this second interpretation of Prigogine’s meaning (which I owe to the coauthor of Order Out of Chaos, Isabelle Stengers),39 there would be no difference in principle between the positions of Whitehead and Prigogine.  They would agree that the basic meaning of time is rooted in a primitive level of experience, which it is the task of metaphysics to articulate.  The differences would be due solely to the fact that, whereas Whitehead moved explicitly into a metaphysical or ontological context, Prigogine has sought to remain within the constraints of science, in which the meaning of all terms must be defined operationally (without confusing this operational meaning with an ontological one).

Whichever interpretation be the correct one, Prigogine’s work is extremely significant from the viewpoint of process philosophy.  For one thing, under either interpretation it goes a long way toward overcoming the dichotomy between the “two cultures,” insofar as this dichotomy has been based on the assumption that physics, as the basic natural science, is essentially nontemporal.  Also, Prigogine’s work provides some verification of the Whiteheadian hypothesis, for if this hypothesis is correct—that cumulative and hence irreversible processes constitute the very nature of actuality as such, and hence of the processes at every level of our particular world—then the more physics advances, the more evidence of irreversible processes it should find.  That is, the metaphysical or ontological irreversibility pointed to by Whitehead should give rise to precisely the kind of empirically detectable irreversibility focused upon by Prigogine.

If the second interpretation of Prigogine is accepted, then his position and Whitehead’s are complementary, as physics and metaphysics, each providing what the other cannot.  But if the former interpretation of Prigogine be accepted, then there is a basic philosophical difference:  whether “time’s arrow” is to be regarded as rooted in a kind of process that is irreversible not only contingently, and hence whether at the deepest level of the world irreversibility reigns supreme.  This could be regarded as the issue as to whether the very meaning of time is to be provided by metaphysics, as Whitehead holds, or whether physics begins with a metaphysical notion of time but then has the task of improving upon it, as Prigogine sometimes seems to say (1973, 590).

However, there is another way of viewing the fundamental issue.  Instead of a contrast between science and metaphysics, a contrast between a modern and a post-modern science could be entertained.  From a post-modern perspective one could challenge the modern assumption that science cannot speak of subjectivity or experience, that the human soul and its analogues must be left outside of natural science as if they were, as in Descartes’s dualism, outside of nature itself.  Whitehead himself was ambivalent on this subject.  He often spoke as if natural science had to remain with the categories of objectivity (see note 31, above); his protest against the fallacy of misplaced concreteness was a reminder that, to get at the fuller truth, we must include the scientific description of nature within a fuller metaphysical account (e.g., 1966, 18, 156).  However, Whitehead also said that science’s categories are not irreformable, and that he was raising his protest on behalf of science itself (1926, 121, 122, 128).  In this latter mood he seemed to be saying that a more inclusive science could include categories of subjectivity, such as “experience,” “concrescense,” and “prehension.”  Prigogine’s drive is certainly to find time rooted more and more deeply in the nature of reality.  For example, he has said:  “It would be quite appealing if the atoms’ interaction with photons (or unstable elementary particles) already carried the arrow of time that expresses the global evolution of nature” (Prigogine and Stengers, 288).  Whether he will move on to a physics that is more fully post-modern or post-classical, rooting time in something like Whitehead’s “concrescence” or Bohm’s “enfoldment,” remains to be seen.

 

1.4. Physics and Pantemporalism

Thus far I have written as if physics could be perfectly compatible with a pantemporalist perspective.  But is this so?  What of the arguments of those who claim the opposite—that physics entails the rejection of the view of time outlined above, with its closed past, open future, and creative present?  For example, Willard Quine has said that the principle of relativity “leaves no reasonable alternative to treating time as space-like” (Quine, 1960, 172). Costa de Beauregard has said:

There can no longer be any objective and essential (that is, not arbitrary) division of space-time between “events which have already occurred” and “events which have not yet occurred.” . . . Relativity is a theory in which everything is “written” and where change is only relative to the perceptual mode of living beings.39

P. C. W. Davies states that:

relativity physics has shifted the moving present out from the superstructure of the universe, into the minds of human beings, where it belongs. . . . The four-dimensional space-time of physics makes no provision whatever for either a “present moment” or a “movement” of time. (Davies, 1976, 2f., 21)

Davies then quotes with approval Herman Weyl’s statement that “the objective world simply is, does not happen.”  Fritjof Capra says:

The relativistic theory of particle interactions shows thus a complete symmetry with regard to the direction of time. . . . This, then, is the full meaning of space-time in relativistic physics.  Space and time are fully equivalent. . . . To get the right feeling for the relativistic world of particles, we must “forget the lapse of time.” (Capra, 1975, 183-85)

A clearer example of the spatialization of time could not be wanted.  First time and space are said to be equivalent; then this equivalence is taken to mean that time is eliminated.  Capra then quotes with approval the famous statement of de Broglie (which is included among the epigraphs at the head of this essay).

However, it appears that the fallacy of misplaced concreteness has been committed again.  The abstractness of the time in the abstract space-time of physical theory is forgotten, and it is treated as if it could be equated with the temporal structure of reality itself.  As Whitehead says, “an abstraction is nothing else than the omission of part of the truth” (1966, 138).  The same point was seen clearly by Mary F. Cleugh.  In Time and Its Importance in Modern Thought, published in 1937, and still one of the best books on the topic, she asks the crucial question:

A fundamental feature of time as experienced is its irreversibility: is this really so, or is it merely an anthropomorphic prejudice, and is physics right in abstracting from this? (Cleugh, 1937, 49)

Her answer is that we need to distinguish between legitimate and falsifying abstraction and that the physicist’s abstraction from time’s irreversibility becomes a falsifying abstraction when it is taken to be a metaphysical truth:

It cannot be too often emphasized that physics is concerned with the measurement of time, rather than with the essentially metaphysical question as to its nature. . . . We must not believe that physical theories can ultimately solve the metaphysical problems that time raises, or that they have any special relevance to these problems. (51)

In support of the abstractness of the physicist’s time, she quotes A. A. Merrill’s article, “The t of Physics”:

t, while created originally from our direct experience with real time, is subsequently handled in a way that has no relation to real time at all.40

She then concludes:

If it is claimed, whether openly or by implication, that the characteristics of “t” give a finally satisfactory account of time, that claim is unfounded. . . . The pliable, reversible “t” may be very useful and important in its own sphere, but its sphere is not that of metaphysics. (Cleugh, 1937, 50)

(Lest some be tempted to consider this type of point irrelevant, on the grounds that they are not interested in metaphysics, it should be emphasized that any attempt to state something about the true nature of time is “metaphysical,” as the term is being used here.)

In an article entitled “Time Represented as Space,” Nathaniel Lawrence argues similarly, addressing “the vanity which holds that the abstract considerations of material science provide an adequate framework for understanding our experience of temporality” (Lawrence, 1971, 123f.).  Physics, he says, is a mode of practice focused upon measurement.  For this purpose, it abstracts from time’s passage, its additive (or cumulative) character, its absolute difference from spatiality, and its qualitative aspects.  The result is time represented as space.  There is justification for this abstract representation in the need for measurement.  But there is also danger, especially because in certain respects physics has been so successful:

The great danger in these restricted enterprises is success.  Success in one’s own particular practice convinces him that he has got his hands on the primary reality.  And therefore the more he will argue that other visions of reality are best tested by one’s own particular discipline. (123)

But the truth, Lawrence says, is that

There is no mode of practice whose presuppositions are adequate for generating a total philosophy. . . . Measurement is almost as hopelessly partial as an approach to reality as is the marketing of peas. (123, 129)

Lawrence’s position reflects that of Whitehead (1958, 27; 1958, 11), who stressed that we must distinguish between the authority of science to establish its own methodology and its competence (i.e., lack of it) to establish our ultimate categories of explanation.

In the same vein is the thought of Arthur Eddington, who points out that the abstractness of time in physics is not unique to it.

Physics has no concern with the feeling of “becoming” which we regard as inherently belonging to the nature of time, and it treats time merely as a symbol; but equally matter and all else in the physical world have been reduced to a shadowy symbolism. (Eddington, 1968, 22)

Like Whitehead, Eddington is trying to warn people not to equate these shadowy symbols, or abstractions, with the concrete realities of the world.  With reference to relativity theory in particular, Eddington says:

Those who suspect that Einstein’s theory is playing unjustifiable tricks with time should realize that it leaves entirely untouched that time succession of which we have intuitive knowledge, and confines itself to overhauling the artificial scheme of time which Römer first introduced into physics. (Eddington, 1968, 18)

Mary Cleugh comments on the widespread idea that relativity reduced time to a dimension of space.

In the new physics, when it is said that the dimensions of “space-time” are at right angles to each other, that does not mean . . . that time is somehow reduced to a dimension of space.  It means precisely the contrary. (Cleugh, 1937, 69)

For support, she cites Einstein himself:

The non-divisibility of the four-dimensional continuum of events does not at all, however, involve the equivalence of the space co-ordinates with the time co-ordinate.  On the contrary, we must remember that the time co-ordinate is defined physically wholly differently from the space co-ordinates. (Einstein, 1950, 31)

Likewise, in reply to Emile Meyerson’s complaint against the tendency of many to reduce time to a fourth dimension of space (see introductory epigraph), Einstein wrote:

Meyerson rightly insists on the error of many expositions of Relativity which refer to the “spatialization of time.” . . . The tendency he denounces, although often latent, is nonetheless real and profound in the mind of the physicist, as is unequivocally shown by the extravagances of the vulgarizers and even of many scientists in their expositions of Relativity.41

Phillip Frank also points out that the metaphysical “extravagances” regarding relativity theory are not only due to the “vulgarizers,” but “have their origin in the insufficiently clear formulations which can be found in treatises of physics themselves” (Frank, 1976, 388).  In the article entitled “Is the Future Already Here?”  Frank’s target is the attempt to justify from relativity theory the metaphysical view “that everything that happens is determined from all eternity, and that there is no development and nothing really new in the world” (387).  As an example of an extravagant metaphysical claim based upon confusion, he cites (389) James Jeans’s statement:

It is meaningless to speak of the facts which are apt to come . . . and it is futile to speak of trying to alter them, because, although they may be yet to come for us, they may already have come for others.42

Milič Čapek has devoted the most attention to this topic in recent times.  He rejects as confused those inferences from relativity theory according to which the relativization of simultaneity destroys the objectivity of temporal order, so that events succeeding each other in one inertial system could appear in reverse order in another system (Čapek, 1976, 506, 507, 511).  The truth is, he argues, that there is much absoluteness in relativity theory:

The irreversibility of the world lines has an absolute significance, independent of conventional choice of the system of reference. . . . Since the universe consists of the dynamical network of the irreversible causal lines, their irreversibility which remains absolute in the relativity theory is conferred to the universe as a whole. (514, 515)

Since (the causal) “before-after” relation is invariant in all systems, it follows that in no frame of reference can my particular “here-now” appear simultaneous with any event of my causal future or with any event in my causal past. (518f.)

The idea expressed in the above quotation from Jeans is categorically rejected by Čapek:

No event of my causal future can ever be contained in the causal past of any conceivably real observer. . . . No event which has not yet happened in my present “here-now” system could possibly have happened in any other system. (519)

Thus the virtualities of our future history which our earthly “now” separates from our causal past remain potentialities for all contemporary observers in the universe.  Something which did not yet happen for us could not have happened “elsewhere” in the universe. (521)

Hence, whether one sees the “time” of relativity theory to be too abstract to be of direct relevance for the philosophical discussion of time, with Cleugh, Eddington, and Lawrence, or whether one sees relativity theory as having philosophical significance, with Čapek and Prigogine, there is good reason to question the counterintuitive ideas that some physicists, speaking as metaphysicians, have claimed to be warranted by it.

What about the claim that, according to elementary particle physics, particles can move backwards in time?  This notion has come into circulation largely through the influence of Richard Feynman’s diagrams and suggestions.  Feynman suggested that, rather than interpreting certain interactions as the production and annihilation of particles, it is simpler and preferable to speak of some of the particles as going backwards in time.  However, P. J. Zwart has argued convincingly that the original interpretation in terms of pair production and pair annihilation is much less problematic than the interpretation based on time reversal (Zwart, 1976, 155-59).  Insofar as the idea of the production and annihilation of particles is thought to be problematic, since it suggests creation out of nothing, Milič Čapek comment (1976, 517) is relevant:

It is not true that the process described above involves “creation from nothing” and “vanishing into nothing”; the pair of particles arises from electromagnetic radiation, into which it can be reconverted.

As pointed out earlier, Whitehead’s suggestion that a particle is really a temporally-ordered society of momentary energy events makes the idea of the emergence and disappearance of “particles” from a background field of relatively chaotic energy events seem less counterintuitive than it would otherwise be.

Hence, there is no good reason to hold that any conventions in physics should lead us to think that time is unreal, or reversible, on the grounds that there are things that really “move backwards in time.”  Even if we had no other reason for rejecting this suggestion, the mere fact that no one can say what that conjunction of words means should give us pause.

I close this section by pointing out that Prigogine emphatically rejects the view that quantum and relativity theory should give any aid and comfort to eternalists, even if they could have once been so interpreted.  As he sees it, quantum theory now describes the transformation of unstable particles, and general relativity is now seen to describe the thermal history of the universe (Prigogine and Stengers, 1984, 9).  Although Einstein’s own 1917 picture of general relativity presented a static, timeless, Spinozistic vision of the universe, it was soon shown that there are time-dependent solutions to his cosmological equations (215).  Finally, instead of regarding the joining of space and time in relativity as implying the spatialization of time, Prigogine agrees with those, such as Bergson and Čapek (1983), who say that it is more accurate to speak of the temporalization of space (17).

 

1.5. Importance of the Topic

Why is the topic of this volume important?  That is, why is it important whether time, with its asymmetry, becoming, and irreversibility, is ultimately real or illusory?  And why is it important to get clear about the relevance of physics to time and of time to physics?  I shall conclude this introductory essay with some reflections on these questions—still from a perspective that is deeply shaped by process philosophy.  I shall organize this discussion in terms of six reasons for the relevance of the topic.

(I) The importance of a temporalized physics for overcoming the dichotomy that has existed in the modern world between the natural sciences and the humanities has already been discussed in relation to Prigogine’s thought; it need not be elaborated upon further.

(2) However, there is also the problem of the dichotomy within the natural sciences, between fundamental physics and everything else.  For, as Stephen Toulmin and June Goodfield (1965, 247) point out in The Discovery of Time,

The physical sciences had stood aside from the historical revolution which transformed the rest of natural science, taking it as axiomatic that certain aspects of the world remained fixed and permanent throughout all other natural changes; and though. . . the list of these timeless entities. . . is much shorter than it was in 1700, the existence of unchanging physical laws, at least, is still regarded as one enduring aspect of the natural world.

Like Prigogine, Toulmin and Goodfield see that this idea of physics was conditioned by theological ideas:  “since God Himself was regarded as changeless and eternal, it was presumed that the ‘laws of nature’ which were an expression of His will were correspondingly fixed in their form” (264).  They see it to be “still an open question whether physics will ever become a completely historical science” (247).

The outstanding question now, is whether the laws of Nature themselves—the last a-historical feature of the physicists’ world-picture—will in their turn prove to be subject to the flux of time. (250)

This idea had been developed by Whitehead, who suggested that we look upon the so-called laws of nature as simply the most long-lasting and widespread “habits of nature,” i.e., the habits taken by the most elementary processes constituting our particular universe (1938, 154).  This means that “natural laws” and “sociological laws” would be different only in degree, not in kind.  They would be more or less universal patterns of behavior which tend to be transmitted from generation to generation, but which can also undergo more or less gradual changes.  In Whitehead’s view the so-called elementary particles, such as electrons, protons, and neutrons, which emerged at some point in the past (1958, 24f.), are really temporally-ordered societies of momentary energetic events that have been stable enough to endure for billions of years.  They did not emerge from absolute nothingness, but from a realm of chaotic events.  Surely it is hard to think of the laws that describe the behavior of all such “particles” as preexisting them.  Accordingly, if the idea that the “elementary particles” of nature have evolved is accepted, it is most natural to think of the laws of physics as themselves having developed at some period in the past.  And, once this thought is accepted, it is natural to assume that their validity will be for a limited duration (1926, 65), i.e., that they will evolve some time in the future, when the entities whose behavior they describe evolve.

The idea that electrons, protons, and neutrons have evolved historically from a relatively chaotic field of events has been given additional support by more recent developments in elementary-particle physics, as now there are considered to be over a hundred forms of such “particles.”  But most of these forms exist only very briefly, many for less than a billionth of a second.  This constant appearance of other particles from the background field and disappearance back into it supports the notion that electrons, protons, and neutrons also emerged from such a background and are simply much more stable forms of temporally-ordered societies of momentary energetic events—but not so stable that they will not eventually disappear.

Also, we have learned that the various forms of elementary particles, including photons, can be transformed into other types.  This further supports the notion that all the basic forms of matter are only partially stable “societies” of momentary events, which have evolved into their present forms and which may presently be undergoing some slow and subtle evolution.  Again, if what we call “matter” has itself evolved, is still evolving, and will some day no longer exist, how can it make sense to speak of the “laws of matter” as themselves timeless?  Must we not assume that these laws evolved correlatively with the entities whose behavior they describe, just as biological laws evolved along with the living things whose behavior they describe and as sociological laws evolved with the types of human beings whose behavior they describe?  If this idea is accepted, then time, with its difference between past, present, and future, its “moving now,” and its irreversibility in principle, will have to be recognized as applying to the fundamental laws of physics, for it will thereby be recognized that the so-called fundamental laws themselves have a history, that their importance in the actual world had a beginning and will have an end.43

When this occurs, then physics will finally, with the rest of the physical sciences, have discovered time, and one more absolute dichotomy which prevents our developing a unified world view on the basis of the evolutionary paradigm will have been overcome.44

(3) The dichotomy between temporal and nontemporal is correlative with a dichotomy between freedom as real and as illusory, for without becoming, in which the present involves turning potentiality into actuality, there can be no freedom (as stressed in Frederick Ferré’s concluding essay).  This dichotomy has led to tragic divisions within human experience.  On the one hand, the natural sciences in general, and physics in particular, have gained such authority in our culture that it is very difficult to believe something wholeheartedly if physical theory seems to contradict it.  On the other hand, it is also impossible, as I have suggested, consistently to deny those “hard-core commonsense” notions which all people presuppose in practice, as evidenced by their behavior.  “Freedom,” the idea that we are to some degree free to determine our own responses to events, is one of those notions.  It implies the partial openness of the future, and yet many have taken physics to imply that the “future” is already as fully determinate and actual as the past; see, for example, the epigraphs at the head of this essay by de Broglie, Weyl, Grünbaum, and Einstein.  Hence, people have been torn between contradictory ideas.  Einstein himself proves an example.  In one of the epigraphs, he says that freedom is an illusion, but then adds tellingly:  “even if a stubborn one.”  His own life displays a marked contrast with his writings.  In his writings he portrays a Spinozistic universe, in which all things are eternally determined (Einstein, 1954, 48-50, 55-56); but he devoted much of his energy to the passionate quest to avert atomic war-showing by his practice that he knew that our fate is not already in the stars (121-67).  Finally, Rudolph Carnap reports that, near the end of his life, Einstein was deeply worried by the awareness that there is something about the present Now that makes it essentially different from the past and the future, but that this is a difference which “does not and cannot occur within physics” (Carnap, 1963, 37f). (For more on Einstein’s views, see Popper, 1982, II, 2-3n, 89-92.)

(4) The tension between fate and freedom is only one aspect of the dichotomy between the sciences and the humanities.  As Prigogine stresses, the dichotomy of the temporal and nontemporal is at the root of a more general opposition between humanity and nature, which involves the clash between the qualitative and the purely quantitative, between the active and the passive.  The idea that the world of nature is a passive realm, devoid of aesthetic qualities and intrinsic value, has led to a devaluation of these values in modern life.  Hence, the issue of the ultimate status of time is part of the overall problem of modern thought, which has, among other things, contributed significantly to the ecological crisis of our time.

(5) Another issue is the tension between modern physics and traditional religions.  In much semipopular writing, the impression is given that traditional Christian thought has affirmed the asymmetry, and hence ultimate reality, of time, whereas modern physics gives us this great, new, liberating idea that the relations of the present to the past and to the future are symmetrical, so that time is unreal; and that, on this point, modern physics agrees with Buddhism and other forms of mystical and perennial philosophy.

However, as the song puts it:  “It ain’t necessarily so.”  On the one hand, traditional Christian theology has portrayed God as eternally and immutably knowing the world.  This means that what is future to us has to be as present to God as what is past and present to us.  Indeed, God was said to know the entire history of the world in one simultaneous now (simul nunc).  It was only with recent process philosophies and theologies that we have Christian theological systems that genuinely allow for the reality of time.  It is the reality of time, not its denial, that is the new idea in Western philosophical and theological thought.44a  

On the other hand, Buddhism by no means unambiguously affirms the symmetry of past and future.  The mutual interfusion of past, present, and future, which Fritjof Capra stresses in The Tao of Physics, was mainly advocated by one school of Chinese Buddhism, Hua-Yen,45 which then influenced some branches of Japanese Buddhism, including Zen.46  There are Buddhists who strongly disagree with the idea of temporal symmetry and who see it as antithetical to Buddhist religion.47  For one thing, this idea that the future is as influential upon the present as is the past stands in strong tension with the idea of karma, which pervades all Buddhist thought, according to which the causal influences upon one arise from the past, but that one can so act in the present as to become liberated from bad karmic influences.

Finally, Ken Wilber, one of the leading theoreticians of transpersonal psychology, which is based on mysticism and the so-called perennial philosophy, emphatically rejects the idea that the relation of the present to the future is no different from that to the past.  In fact, Wilber endorses Whitehead’s views on this subject: an event prehends only its predecessors, not its descendants; Christopher Columbus affects us, but we do not affect him (Wilber, 1982, 286-88).  Evidently neither Buddhism, nor mystical experience, nor the “perennial philosophy” can be appealed to as an unambiguous witness to the mutual interpenetration of past, present, and future and hence to the ultimate unreality of temporal distinctions.  Just as it is not a question of what “physics” says, but of what individual physicists or philosophers of physics say (as pointed out in the introductory epigraph by Gale), so it is not a simple question of what “mystical experience” or “enlightenment” or “Buddhism” says, but of what is said by individual interpreters, with their particular histories, biases, preconceptions, and selective attentions.  As Whitehead says, if you want uninterpreted experience, ask a stone for its autobiography (1978, 15).

Wilber also makes reference to so-called precognitive experiences, indicating that, however they are to be interpreted, they must not be interpreted so as to deny the reality of free will in the present (Wilber, 1982, 287).  This is a point upon which I want to enlarge.  We could probably not exaggerate the extent to which ideas of the ultimate unreality of time in religious thought have been based on such experiences.  Such experiences—in which a vision turns out to correspond with a later event—have been widespread, especially in contexts out of which religious writings arose.  The most common interpretations of these experiences have been these: (a) God reveals to a prophet what God “already” knows is going to happen; (b) in meditation, one comes into contact with a level of existence in which all things, which in ordinary consciousness seem to be distinguishable as past, present, and future, exist simultaneously in an “eternal now”; or (c) a future event causes the present vision.  All three of these interpretations imply that that realm which appears to be future from the standpoint of any present “now” is in reality as filled with fully determinate events as is the past.  Hence, there is no categorical difference between past and future, so it is purely arbitrary, as Bertrand Russell says, that we remember only the past and not the future.  So-called precognitive experiences are widely taken as evidence that this arbitrariness is sometimes overcome, as some people “remember” events in the future.

However, these are all tendentious interpretations.  Purely phenomenologically, what is experienced is (i) a vision followed by (ii) an event that corresponds closely to the vision (too closely to be dismissed as mere coincidence).  That is, there is nothing in the experience as such that dictates that when the vision occurred the event already existed or that the (later) event caused the (prior) vision.  It would be equally compatible with the evidence to say that the vision caused the subsequent event to occur.  In fact, there are thoughtful interpretations of so-called precognitive experiences that employ this notion (Eisenbud, 1956, 23-25; Ebon, 1968, 224; Tanagras, 1967).  However, most so-called precognitive experiences can be explained within a pantemporalist framework without resort to such a drastic idea.  For example, most so-called precognitions of disasters can be accounted for in terms of unconscious clairvoyant knowledge of present structural defects (e.g., a leaky ship, a cracked wing, a short circuit in electric wiring, a weak heart, a cancerous growth), plus unconscious inference of the probable outcome, resulting in a dreamlike vision rising to conscious awareness.

Most other so-called48 precognitions can be accounted for by unconscious telepathic knowledge (of other people’s knowledge, feelings, or intentions—perhaps unconscious on their part), plus unconscious inference.  For other cases not easily explainable in one of these three ways, there are several other ways that do not require reverse causation, or an already actual future, or anything else implying the unreality of time.49  So, again, there is nothing about these special experiences that should encourage us to deny the temporality of reality, which is suggested by the overwhelming majority of our experience, secular and religious alike.

To sum up the main point of this discussion: in response to the claim of Prigogine and process philosophers that a nontemporal interpretation of physics isolates it from the rest of human culture, some might reply, “Not entirely.  Certainly, it isolates physics from conventional interpretations of experience and reality, including conventional Western religious thought, but it puts physics in harmony with those experiences in which human beings have transcended conventional consciousness and plunged more deeply into the nature of reality.  On the basis of those experiences have arisen interpretations of reality that are remarkably similar to those arising from modern quantum and relativity physics.”  Over against this view, my argument has been that a nontemporal view of reality is not unambiguously supported by any kind of nonscientific experience—ordinary, religious, or parapsychological.  Accordingly, there is good reason to hold to the original point:  nontemporal interpretations of physics will continue to stand in contradiction with the rest of human experience and the best interpretations of reality based on it.

(6) There is special reason for focusing upon the question of physics and the ultimate significance of time at the present moment.  There has recently been a spate of popular interpretations of physics that have reached wide audiences.  These books have spread quite widely the idea that the authority of physics supports the notion that time is unreal.  Reference has already been made to Fritjof Capra’s immensely popular Tao of Physics.  Gary Zukav’s Dancing Wu Li Masters has also been a best-seller.  In it, the reader learns that the preferred interpretation of quantum field theory is to speak of anti-particles as particles traveling backwards in time (Zukav, 1979, 236, 237).  The reader encounters a quotation from de Broglie and learns that it is an illusion that events “develop” in time (238).  Zukav’s moral is similar to Capra’s: since the flow of time has no meaning at the quantum level, and consciousness itself may be a quantum process, we can perhaps experience timelessness (240).  Fred Alan Wolf, whose Taking the Quantum Leap was quite popular, now tells us in Star Wave not only that time is unreal (1984, vii, 230), but that it is in our power to create the past (vii, 109, 110, 187), so that what “really happened” is determined by a majority vote (230, 326).  Accordingly, it is within our power to determine whether the Nazi Holocaust occurred—not just epistemically, but ontically: we could make it so that it really had not happened.50

These metaphysical specul’ations, which claim to be warranted by the “new physics,” are particularly egregious examples of the havoc that can be created in people’s thinking by the fallacy of misplaced concreteness.  The interpretation of quantum physics from which this conclusion of the unreality of time is drawn is one which methodologically abstracts from the question of what is really going on in the microscopic realm.  It rejects the attempt at a realistic interpretation in favor of a phenomenalistic one.  That is, it restricts itself to formulae that successfully correlate the macroscopic observations, i.e., the instrument readings.  (Bohm’s long interest in “hidden variables” was less motivated by the desire to confirm Einstein’s and de Broglie’s deterministic vision than to find a theory that would describe the real processes going on.)  But then interpreters forget the abstraction from reality and start applying details of the intentionally phenomenalistic account to the operations of the real world.  In Wolfs case, the crucial idea is that no quantum reality has determinate status until it is measured, it being the measurement that is said to “collapse the wave function.”  The paradoxical implication, drawn unflinchingiy by some interpreters, is that there can be no determinate reality apart from human observers, so that even what we normally think of as the distant past is dependent upon present observation.51  This would mean that the entire theory of evolution would have to be abandoned, for that theory (which virtually all of contemporary science presupposes) assumes that quite determinate processes did occur for billions of years prior to the rise of human (or humanlike) observers.  Such a conclusion should alert us that something is amiss somewhere.

However, there is little danger that masses of laypeople are going to be led to believe that the past is under their control and hence to begin taking their decisions less seriously on the grounds that, if they do not like the results (e.g., a nuclear war) they can simply revise the past so that those events did not occur (assuming that there are any of us left to carry out the revision).  Our sense of the immutability of the past is too strong to be shaken, even if we be told that we are rejecting the authority of physics.  After all, classical theists, never shy about affirming divine omnipotence, would not even ascribe to God the power to undo the past.  Accordingly, affirming the symmetry of past and future52 usually means regarding not the past as open but the future as closed.  The introductory epigraphs by de Broglie, Weyl, Gold, Griinbaum, Russell, and Einstein all reflect this view.  The effect these ideas can have upon the intelligent layman, even with some scientific training, is illustrated by a book by Larry Dossey, M.D., entitled Space, Time and Medicine.

Dossey begins his chapter on “Modern Time” with the introductory epigraphs by Gold and de Broglie, after another by Bertrand Russell, which reads:

A truer image of the world . . . is obtained by picturing things as entering into the stream of time from an eternal world outside, rather than from a view which regards time as the devouring tyrant of all that is.  (Russell, 1917, 21)

After introducing the idea of creativity, Dossey draws these conclusions:

This concept of creativity as a feature of a timeless, eternal world is hard to swallow, especially if one wishes to see his creative act as a literal making of some new thing.  But in the modern context of a nonlinear time, as de Broglie states, events exist prior to us in time.  We create nothing, since all things exist already. (Dossey, 1982, 34)

As the beginning of that statement shows, Dossey is keenly aware that this view of time goes directly counter to our ordinary notion of time—what I have called our “hard-core commonsense” view.  But, illustrating my earlier point about the authority generally granted to physicists in our culture, he shows no hesitation as to which view to adopt:  “The view of time from modern physics tells us our ordinary notions of time are wrong” (41).  After quoting the famous statement from Weyl, he says:  “This view is an affront to common sense. . . . [But we] must assimilate it,” for “we cannot ignore what modern physical science has revealed to us about the nature of time” (152, 153).  He justifies this rejection of “common sense” by quoting Einstein’s statement that “common sense is merely the deposit of prejudices laid down in the human mind before the age of eighteen.”  This is probably true of most of those notions that are usually classed as “common sense” and that I have called “soft-core common sense.”  But is there not an important difference between those culturally conditioned ideas, out of which we can be educated, and the true (i.e., hard-core) commonsense ideas which belong to the sensus communis of humanity and which we all inevitably presuppose in practice?  Do we not have good reason to have more confidence in them than in any ideas from which their contraries can be deduced?  And does it not belong to our hard-core common sense that the future is not totally determined?  If so, then we should take those ideas as the criteria for evaluating the alleged “revelations” from modern physics, instead of the other way around.

The power of the physicalist conception, that if something cannot be detected by the methods of science then it does not exist, is shown by the fact that Dossey repeatedly cites with approval the statement of P. C. W. Davies that no physical experiment has ever been performed that detects the passage of time (Dossey, 151).  Dossey then provides a summary of an “incontrovertible lesson” from modern physics, which amounts to a complete endorsement of Davies’s views:

The notion that time flows in a one-way fashion is a property of our consciousness.  It is a subjective phenomenon and is a property that simply cannot be demonstrated in the natural world.  This is an incontrovertible lesson from modern science. . . . A flowing time belongs to our mind, not to nature.  We serially perceive events that simply “are”. . . . (151)

Dossey’s concerns are practical.  He believes that this view can help people overcome the anxiety, and related health problems, caused by looking upon time as “the devouring tyrant” (146).  He suggests that we learn from modern physics that time is a chimera and that “length of life is meaningless for the reason that passage of linear time does not occur” (145).  His conclusions are well-intentioned, but whereas it may be healthy, in some sense, to apply to oneself the idea that “length of life is meaningless,” to take this as a generally true statement, and to apply it to one’s enemies, or to humanity as a whole, could be extremely dangerous—especially in this nuclear age.

So, if time is real, and if each moment is an opportunity to decide which among a variety of potential values will become actual, thereby conditioning the entire future of the universe for good or for ill, then the idea that time is not real in this sense is a false and dangerous one.  It will lead us to take our enormous power for decisionmaking, and hence value-realization, less seriously than we should.  The same is true for the idea that the past can be undone.

Given the tremendous authority granted to physics and hence physicists in our culture, and the importance that our ideas about time have in our attitudes toward life and the way we live it, it is crucial that we become clearer about what contemporary physics does and does not entail about the nature and ultimate significance of time.  It is my hope that this volume will make a helpful contribution toward this end.  In this introductory essay, I have pointed out various paradoxes about time that have plagued human thought, especially in the modern period.  I have sought to give reasons for taking seriously in each case the suggestion of Whitehead that “the paradox only arises because we have mistaken our abstraction[s] for concrete realities” (1926, 80f.)

Notes

1 This is the behavior of the K meson; see Whitrow (1980, 355f.); Davies (1976, 175f.).

2 Some writers unfortunately use the term asymmetry for anisotropy.  I use asymmetry to mean that the present’s relation to the past is categorically different from its relation to the future.  When P. C. W. Davies speaks of “the physics of time asymmetry” he does not have asymmetry in this sense in mind, but only what others (e.g., Grünbaum, 1967) less ambiguously call “anisotropy.”  For example, Davies is at pains to distinguish “the real, physical asymmetry” from “the controversial phenomenon of psychological time,” which divides time into the past and the future separated by a now (Davies, 1976, 3, 20-22).

3 “Nothing yet discovered in nature requires individual atoms to experience time asymmetry, the very essence of which is the collective quality of complex systems, like life itself’ (Davies, 1976, 4).

4 Richard Feynman says that “irreversibility is caused by the general accidents of life. . . . It is not against the laws of physics that the molecules bounce around so that they separate.  It is just unlikely.  It would never happen in a million years. . . . Things are irreversible only in a sense that going one way is likely, but going the other way, although it is possible and is according to the laws of physics, would not happen in a million years” (Feynman, 1965, 112).  Writers who so derive the irreversibility of time from entropy hold that our memory is irreversible only because we are ourselves complex systems exemplifying entropic principles (Feynman, 1965, 121; Davies, 1976, 19-22).  Accordingly, if thermodynamic systems were to reverse directions, our memory would presumably reverse too, with the result that we would remember the future but only predict the past.  One should consider this conclusion a reductio ad absurdum of the position.  Percy Bridgman agrees (see epigraph at head of this introduction; Bridgman, 1955, 251); so does P. J. Zwart, who considers it absurd to think the temporal order is based on the entropic order (Zwart, 1973, 144).  He says: “One thing is quite inconceivable: that we could perceive a later event before an earlier one. . . . This kind of proposition is self-contradictory, that is to say, time reversal in this sense is logically impossible.  Reversal of the entropic ordering is not logically impossible, however, though highly unlikely.  But even if it occurred, we should only have to adjust our laws, not our concepts” (Zwart, 1973, 145).  Likewise K. G. Denbigh:  “Mental processes display irreversibility of a kind not shown by physical processes—that is, in the sense that it is not conceivable that they could ever occur in the reverse temporal sequence” (Denbigh, 1975, 39).  He points out that, if we had settled upon some physical process as a standard for deciding the before-after relationship—e.g., an apple falling from a tree to the ground—and then found that the event sequence went in reverse order, we would simply decide that the “standard” was fallacious.  “In other words, is it not the case that what we really mean by the before-after relation is the relation as it is offered by consciousness?” (Denbigh, 1975,41).  My only difference with Denbigh here is that, from the viewpoint of process philosophy, the notion that “physical processes” are reversible results from conceiving of them at an abstract level and hence in effect conceiving of them as types of processes.  If we took experiences, or what he calls “mental processes,” at the same level of abstractness, they also would be considered reversible.

5 Richard M. Gale (1968, 107) says that it is an “analytic truth that a present cause cannot have a past effect.”  Process philosophers agree, while rooting this truth of language in the structure of experience and finally of actuality in general.

6 Even Einstein evidently came to believe this; see the discussion in section 1.4 in the text below.

7 See, for example, H. Bondi (1952) and H. Reichenbach (1956, 269).

8 Henry Mehlberg (1980, 202) asks how we can reconcile the fact that our knowledge of the past differs so greatly from our anticipation of the future with the fact, vouchsafed by science, that time is isotropic, having no intrinsic difference between past and future.  He suggests that “the only possible answer is the realization that temporal words like ‘past,’ ‘future,’ ‘planning,’ etc. have no independent meaning,” but depend upon the usage of one’s linguistic community.  “What I remember belongs to my past, by definition, and what I desire or am planning for, belongs to my future, by definition .equally.  But this need not prevent somebody else from desiring what I remember and, thus, from having his future overlapping with my past.”  Should such a conclusion not lead one to suspect that something is amiss?

9 Adolf Grünbaum (1967, 152f.) says that “the transient now with respect to which the distinction between the past and the future of common sense and psychological time acquires meaning has no relevance at all to the time of physical events.”  Accordingly, “coming into being (or ‘becoming’) is not a property of physical events themselves but only of human or other conscious awareness of these events” (I 54).

10 The foregoing account was meant as an explication of those features that are largely agreed upon by authors from diverse perspectives, not as an adequate account of time as we experience it.  For it excludes the crucial feature, discussed below in the explication of Whitehead’s view, that time as experienced always involves conformity to the past (and anticipation of the future’s conformity to the present).  It is this conformity to (and in fact inclusion of) the past that is the strongest reason for the irreversibility of time.  The theory of time including this feature is a causal as well as relational theory.  My one major difference from P. J. Zwart arises here, in that Zwart (1973, 131-33, 144) holds that time is to be understood relationally but not causally, as he regards the temporal order as more fundamental than the causal.  However, the difference seems to arise because he thinks of events, causes, and effects as abstractions, i.e., types of events.  This is clear in his argument that if “we should suddenly perceive that all the events we used to call the effects of other events were preceding these latter events, it is inconceivable that we should adjust the temporal order to the causal order, instead of the reverse” (144).  Apparently for this reason he can say: “If temporal order could be reduced to causal order ‘A precedes B’ would have to imply ‘A is the cause of B’ which, of course, is not at all true” (144).  But if we understand A and B not as types of events but as concrete, singular events, and also that by calling A a “cause” of B we mean that it was an event without which B could not have been precisely what it was, then Zwart could perhaps accept the equation of causal and temporal order.

11 As K. G. Denbigh says, the view that the universe is symmetric temporally would lead to the idea that the “happening” or “taking place” of events must be thought of as a sort of illusion of consciousness that has no objective counterpart (1975, 7).  In the introduction to a symposium on time, Thomas Gold (1967, 2) asks whether the notion of a flow of time, which we get from introspection, is “only a deception of a biological sort which ought to have no place in physics.”  Later he reveals that this is indeed what he holds:  “We ought to eliminate this flow idea from the real picture [of the world], but before we can eliminate it we ought to understand how it arises.  We should understand that there can be a self-consistent set of rules that would give a beast this kind of phoney picture of time” (182).  He closes the volume with a hope for “some new physical theory” that will give a “better description of nature, one that is less dependent on our subjective notions” (243).  P. C. Davies (1976, 22) speaks of “the apparently illusory forward flow of psychological time.”  His reductionistic position is especially revealed in a statement following his observation that, although the laws of physics do not provide a time asymmetry, the world as a matter of fact is asymmetric (i.e., anisotropic;  see note 2) in time.  He concludes from this that the world’s anisotropy is hence “extrinsic” or “fact-like” rather than “intrinsic” or “law-like.”  These equations show that, for him, nothing is intrinsic to the world unless it is contained in the fundamental laws of physics.  My object is not to criticize those who hold these positions but to show, first, that many people have adopted such positions and, second, to suggest a framework in which both their concerns and a more adequate view of time can be combined.

12 See note 4.

13 For example, Mary F. Cleugh (see the first epigraph by her) and Robert S. Brumbaugh (1984, 6,10): “The study of time is a metaphysical enterprise, not a physical one.. . . The nature of time is a philosophic problem.  In physics, time is an undefined basic notion.”

14 Whitehead describes the fallacy as “mistaking the abstract for the concrete” (1926, 74f.)  One major form of it is the description of actual things as if they exemplified “simple location,” i.e., as if they could adequately be described as simply being at one place in space-time, apart from any essential reference to other regions of space and other durations of time (1926, 74f., 84).

15 “The reality is the process” (1926, 106); “the very essence of real actuality—that is, of the completely real—is process (1933, 354).

16 In this volume; and Čapek (1976, xxxiv-xxxv).

17 I have focused on this aspect of process thought in relation to the problem of evil in Griffin (1981).

18 See notes 8 and 11, and the first epigraph by Einstein, above.

19 I have put “elementary” in scare quotes to indicate distance from the notion of “elementarism,” which is discussed in Bjelland’s essay in this volume.  Elementarism implies that there are fundamental particles out of which more complex things are created but which themselves undergo no essential changes by becoming part of this new environment.  This is exactly the notion of autonomous bits of matter with “simple location,” which process philosophy rejects as a fallacy based on misplaced concreteness (see note 14).  The term “elementary particles” also suggests that these entities are ultimate, i.e., that they do not require explanation, and are not thought to have evolved out of something more ultimate.  Process philosophers share with Karl Popper (1982, III, 139, 158) the rejection of this notion of elementary particles.

20 For Whitehead, the “metaphysical rule of evidence” should be “that we must bow to those presumptions which, in despite of criticism, we still employ for the regulation of our lives” (1978, 151).  It is these presumptions, which are inevitably presupposed in practice (Whitehead, 1978, 13), that I call our “hard-core commonsense notions.”  The chief problem with the so-called empiricist philosophies of Locke and Hume and their followers was that they did not take these empirical facts as basic but instead began with certain a priori dogmas about experience from which they either deduced the falsity of these hard-core commonsense notions (13, 146), or at best added them as presuppositions of practice as supplements to their metaphysical or epistemological theory (133f., 153, 156).

21 Although Richard Gale (1968, 103f.) does not explicitly distinguish between hard-core and soft-core commonsense, his position is very similar to mine.  In his words, “many of our fundamental concepts, such as causality, action, deliberation, choice, intention, memory, knowledge, truth, possibility and identification, logically presuppose an asymmetry between the past and future. . . . These logical asymmetries interlock . . . such that one of them cannot be jettisoned without giving up the others as well.  Together these asymmetries form the basis of our common-sense conceptual system.”  Karl Popper also points out that a major reason for rejecting a symmetrical view of past and future is that it conflicts with our common sense, which regards the past as closed but the future as not completely fixed.  He says:  “We should not be swayed by our theories to give up common sense to easily.”  But, failing to distinguish between soft-core and hard-core common sense, he does not consider common sense the ultimate arbiter, and hence would be willing to give it up in the face of a good scientific theory supporting the symmetrical view (1982, II, 3n., 55-56).

22 This is Roderick Chisholm’s way of stating the “commonsense” approach:  “I assume that we should be guided in philosophy by those propositions we all do presuppose in our ordinary activity. . . . Any philosophical theory which is inconsistent with any of these data is prima facie suspect” (Chisholm, 1976, 15, 18).

23 See Thomas Gold’s statement cited in note II.

24 Čapek (1976, xlvii); see also G. J. Whitrow (1980, 350).

25 This would be one of my main reasons for rejecting the view of Robert Brumbaugh (1984, 9, 11, 125, 131, 138, 139), according to which there are four kinds of time in the world.  I agree with the statement of G. J. Whit row with which he distinguished his own position from that of J. T. Fraser (1975), which is similar to Brumbaugh’s:  “at all levels time is essentially the same, although certain aspects of it become increasingly significant the more complex the nature of the particular object or system studied” (Whitrow, 1980, 375).  See also the discussion of Karl Popper’s position in note 30a.

26 See note 14.

27 “[The] Kantian doctrine accepts Hume’s naive presupposition of ‘simple occurrence” . . . for the mere data.  I have elsewhere called it the assumption of ‘simple location,’ by way of applying it to space as well as to time.  I directly deny this doctrine of ‘simple occurrence.’  There is nothing which ‘simply happens.’  Such a belief is the baseless doctrine of time as ‘pure succession.’ . . . Pure succession of time is merely an abstract from the fundamental relationship of conformation” (Whitehead, 1958, 38).

28 Whitehead (1966, 138).  Whitehead uses the term matter to refer to anything with the property of simple location (1926, 72).  He does not think that matter in this sense is a total fiction; rather, he holds that “by a process of constructive abstraction we can arrive at abstractions which are the simply-located bits of material” (1926, 85).  Again, the error is taking the intellect’s abstraction from actuality to correspond, without essential loss, to the concrete actuality itself.

29 To my knowledge, Whitehead never used the term temporally-ordered society.  Instead, he spoke of enduring objects with purely temporal or serial order as having “personal” order (1933, 259, 263; 1978, 35).  However, he himself pointed out that the term person has connotations of consciousness (1978, 35), whereas his societies with personal order need not even be living (1933, 264), and indeed most of them—e.g., electrons, protons, atoms, molecules—are not.  Hence I find the term temporally-ordered society preferable.

30 For criticisms of Grünbaum’s thought on time, see Milič Čapek (1961, 377); and Frederick Ferré (1972).

30a. Karl Popper’s thought provides an interesting counter example to the general correlation, since he is a dualist in regard to the mind-matter issue but not in regard to time.  He agrees with process philosophers that the defense of realism requires the affirmation of the reality of time (1982, II, 3n.), that time’s arrow is not derivative from entropy, that in any case there is more to time’s asymmetry than its arrow, and that this asymmetry is incompatible with determinism (1982, II, 55-56).  He grounds indeterminism and hence the nonderivative reality of time in the notion that all material particles realize and are “propensities”:  each thing is a realization of a prior potentiality and in turn a potentiality for another becoming (1982, III, 205, 207).  This way of affirming the ultimate reality of time, and hence the temporality of physics, is very similar to the Whiteheadian notion that each actual entity is first a partially self-determining actualization of potentialities proffered it by past actual occasions and then in turn provides potentialities for future actualizations.  However, although Popper attributes consciousness to amoebae, and an elementary  memory to crystals and even to DNA molecules, he insists that atoms and subatomic particles are totally devoid of experience (1977, 29, 71).  In his terminology, World I, which consists of purely material entities, existed prior to World II, which consists of psychological qualities (1982, 114-17).  World I objects seem to have ontological as well as temporal priority:  Popper shares “with old-fashioned materialists the view that . . . solid material bodies are the paradigms of reality” (1977, 10; cf. 1982, II, 116, 117).  But this raises two questions:  The first question is the traditional one:  how can non-experiencing matter interact with experiencing mind?  Popper simply ignores this problem, saying that “complete understanding, like complete knowledge, is unlikely to be achieved” (1977, 37).  That is true, but there is an enormous difference between lacking complete understanding of how something occurs and affirming a position that makes the occurrence seem impossible in principle.  The second question is based on the fact that “propensity,” as Popper recognizes (1982, III, 209), is a psychological term.  How are we to think of “solid material bodies” as having “propensities” even though they have nothing even remotely analogous to experience?  The only statement I have found relating to this question is that the propensities are, like forces, “properties not so much of the particles as of the total physical situation; they are, like forces, relational properties” (1982, III, 127).  But I do not see how this helps.  Popper himself would evidently agree, since he says:  “We take indeterminism as a cosmological fact which we do not attempt to explain” (1982, III, 181).  Both of these problems are avoided if we assume that, below cells and DNA molecules, experience simply becomes less and less complex rather than disappearing altogether.  Hence, I maintain my thesis that pantemporalism can be consistently defended only on the basis of panexperientialism.

31 Whitehead’s is an “epochal” theory of time, meaning that time is derivative from the succession of actual events, each of which has a finite (i.e., noninstantaneous) duration.  Although there is a space-time continuum, this is a structure of potentiality.  The temporal process is not in fact continuous; it is not constituted by a series of instants.  For a defense of Whitehead’s position against criticisms, see David Sipfle (1971).

32 G. N. Lewis, who was perhaps the strongest advocate of time symmetry of all time, saw this point clearly.  He saw the importance of conceiving of matter as strictly atomic, in the sense of being made of discrete particles that in no sense entered into each other, for if two things were allowed to diffuse into each other, “such a diffusion would in principle be a phenomenon which by no physical means could be reversed.”  But with discrete particles, the recurrence of every particular state will eventually come about.  See Lewis (1930, 571) and the discussion in Roger Steuwer (1975).

33 See the discussion and quotations in Floyd Matson (1966, 38-56).

34 That H. Feigl’s identism was really a dualism is evident.  After saying that, instead of two types of events, “we have only one reality which is represented in two different conceptual systems,” he makes this qualifying explanation: “on the one hand, that of physics and, on the other hand, where applicable (in my opinion only to an extremely small part of the world) that of phenomenological psychology” (1960, 33; emphasis added).  In other words, all events have physical qualities, but only a few have psychical qualities.  Feigl says this explicitly:  whereas “at most something very much less than a psyche is ascribed to plants or lower animals” (which implies no dualism), “nothing in the least like a psyche is ascribed to lifeless matter” (which does imply dualism).

35 Jacques Monod (1972, 21) has called this the “postulate of objectivity” and proclaimed it to be the hallmark of science itself.  However, this is to dictate either that psychology must be behavioristic, eschewing all references to purpose, feeling, and the like, or that it cannot be a science.  The same would be true of ethology.  Do we really want to define science in such a way that it is forever excluded from speaking about subjectivity?  At first glance, to say that science presupposes the “postulate of objectivity” seems unexceptional: of course science is to be objective!  But this is to assume only one of the two meanings of this ambiguous term.  That is, as long as the term is taken to refer to a method of approach, the postulate is a commonplace.  But the term can also refer to the categories used to describe the objects of study.  Monod takes the postulate of objectivity to include both meanings, with the result that science can use only categories of objectivity (e.g., mass, energy) as opposed to categories applicable to subjectivity (e.g., feeling, purpose).  Science is thereby limited to treating all things as if they were objects in the metaphysical sense, i.e., objects as opposed to subjects.  This limitation is arbitrary and is really nothing but an attempt to enshrine a particular metaphysics (materialistic or dualistic) into the permanent methodology of science.  But, presuming that it is possible to be objective (methodologically) about subjectivity (as phenomenology and nonbehavioristic psychology maintain), should we not define science in such a way that it is able to deal with whatever kinds of things the world contains?  On the question of treating subjectivity objectively, see Schubert Ogden (1966) and Roger W. Sperry (1976).  On modern science’s commitment to the categories of objectivity alone, see Frederick Ferré (1976, chap. 1 and 2).

36 For suggestions as to what a post-modern science might be like, see Ferré (1976, chap. 5) and Stephen Toulmin (1982).

37 All references to Prigogine’s writing in the text, unless otherwise indicated, are to Order Out of Chaos: Man’s New Dialogue with Nature, by Ilya Prigogine and Isabelle Stengers (New York: Bantam Books, 1984).

38 See the discussion near the end of section 1.2, above.

39 Letter to the author, November 1, 1984.

40 Merrill, p. 240.  A similar point, quoted by Emile Meyerson (1976, 353), was made by W. Wien:  “Although the relationship between space and time as revealed by relativity theory is very important, one must always bear in mind . . . that it is not time itself which plays this role, but imaginary time.”

41 Quoted by Čapek (1976, 366f.).  However, as Čapek points out, Einstein was not much interested in this point, and vacillated on it.  For example, the first epigraph from him at the head of this essay suggests that time is a space-like dimension that can be fitted into a geometrization of reality.

42 The quotation is from James Jeans, Man and the Universe (1935).

43 This does not mean that every feature of the present universe is contingent.  Rather, on the assumption that there has always been a plurality of events, in some state or other, it follows that there are some strictly eternal, or metaphysical, principles these events exemplify.  Whitehead calls Process and Reality “an essay in cosmology.”  The task of cosmology is to describe the most general features of our cosmic epoch, while trying to distinguish the truly metaphysical aspects of reality from those aspects which apply only to our cosmic epoch.  See Whitehead (1978, 287-88).

44 See Toulmin,and Goodfield (1965, 247-65).

44a That this fact is not widely understood is suggested by this statement by Karl Popper (1982, II, 5):  “Since St. Augustine, at least, Christian theology has for the most part taught the doctrine of indeterminism; the great exceptions are Luther and Calvin.”  In reality, Luther and Calvin only differed from Augustine, Aquinas and others by being more explicit about their determinism.  See the relevant chapters in Griffin (1976).

45 For an excellent account of the Hua-Yen school, with a critique of its idea of the symmetry of time from a Whiteheadian perspective, see Steve Odin, Process Metaphysics and Hua-yen Buddhism: A Critical Study of Cumulative Penetration vs. Interpenetration (Albany: State University of New York Press, 1982).

46 Capra (1975, 179) quotes this statement from D. T. Suzuki’s explanation of the Zen perspective:  “In this spiritual world there are no time divisions such as the past, present, and future; for they have contracted themselves into a single moment of the present. . . . the past and the future are both rolled up in this present moment of illumination.”  Suzuki goes on to say that “this present moment is not something standing still with all its contents, for it ceaselessly moves on.”  One cannot help wondering, if the past and future (as it actually will be) are already rolled up in the present, what it “moves on” into.  Capra admits this difficulty; in fact, it is this difficulty that provides the rationale for his book, as he adds:  “But modern physics may help” (179).  But there is good reason to doubt that it can.  For one thing, empirical evidence cannot resolve a logical contradiction.  Also, the suggestion of process philosophy is that the aid can mainly go in the other direction.  That is, it is time as known in our experience in its fullness, in which the future is in the present in a different mode than is the past, that can help us understand the reality of time in that aspect of the world physics studies.

47 See, for example, Ryusei Takeda, “Some Reflection on Yogacara Philosophy in the Light of Whitehead’s Metaphysics,” unpublished essay, on file at the Center for Process Studies, 1325 N. College, Claremont, CA 91711.

48 I use the qualifier so-called because the term precognition imposes an unwarranted interpretation upon the experience.  The term suggests that an event can be cognized (known) before it has occurred, which implies that the event in some sense already exists to be known before it has appeared in the present.  As G. J. Whitrow puts it: “unless we live in a ‘block universe’ in which everything in our future is pre-ordained so that we are mere automata completely unable to influence any of our future actions—a situation that seems to conflict with our actual experience—a future event is nothing but an unrealized possibility until it happens and therefore cannot give rise to genuine precognition” (Whitrow, 1980, 367f.).  My only qualification would be that there are objective probabilities, and some so-called precognitive experiences might result from having an intuition of these.  But this would not be genuine precognition.  On the one hand, it would not be cognition of the future event as such, since cognition means knowledge and, so long as an event is merely probable, something could happen to prevent its occurrence; hence, that it is to occur cannot be known in advance.  On the other hand, knowing a probability is knowing something that already exists and hence is not precognition.

49 Of course, many people reject all so-called paranormal occurrences out of hand.  For them, there is no problem to be solved, since there are no precognitive experiences.  Any “solution” that involves an appeal to psychokinesis and extrasensory perception is necessarily as absurd as the imagined problem it means to solve.  However, my point here is only that those who do accept the reality of paranormal experiences in general can hold to the ultimate reality of temporality without arbitrarily denying the genuineness of that class of paranormal events that has led people to speak of “precognition.”  Those forms of paranormal events which do not suggest the causal influence of the future upon the present can adequately account for that one form which does, at first glance, seem to suggest this.

50 This example of the Holocaust was given by Wolf himself in a lecture for a meeting of the World Futures Society held at the University of Southern California on April 28, 1984.  Although the idea of re-creating the past is not strongly featured in Star Wave, in this prepublication lecture, and in private conversation, he lifted up this idea as one of the chief implications of the book.

51 This kind of idea has been around for several decades.  For example, G. N. Lewis (who coined the term proton) suggested that we could prevent light from having been emitted from a star a thousand years ago (Lewis, 1926, 25).  Wolf is dependent primarily upon John Archibald Wheeler, who has defended the idea more recently.  For a popular account, see Wheeler (1982, esp. 14-18).

52 Wolf says he does not affirm that the past and future are symmetrical, but that the past is open while the future is closed, i.e., strictly predestined (I984, vii, 326).  However, in other passages, especially exhortatory ones, it is clear that he, like everyone else, in practice assumes that the future in part is still to be determined.

 

References

Augustine, Saint. 1912. Confessions. Cambridge: Harvard University Press.

Bergson, Henri. 1911. Creative Evolution. New York: Holt.

Bondi, H. 1952. “Relativity and Indeterminacy.” Nature 169, April 19: 666.

Borges, Jorge Luis. 1967. “A New Refutation of Time.” In A Personal Anthology, trans. Anthony Kerrigan, 44-64. New York: Grove Press.

Bridgman, Percy. 1955. Reflections of a Physicist. New York: Philosophical Library.

Brumbaugh, Robert S. 1984. Unreality and Time. Albany: State University of New York Press.

Čapek, Milič. 19M. The Philosophical Impact of Contemporary Physics. New York: Van Nostrand, Reinhold.

_______. 1976. The Concepts of Space and Time. Boston Studies in the Philosophy of Science, Vol. 22. Dordrecht, Holland: Reidel.

_______. 1983. “Time-Space rather than Space-Time,” Diogenes No. 123 (July-September), 30-48.

Capra, Fritjof. 1975. The Tao of Physics: An Exploration of the Parallels between Modern Physics and Eastern Mysticism. Boulder: Shambala.

Carnap, Rudolph. 1963. “Intellectual Autobiography.” In The Philosophy of Rudolph Carnap, ed. P. A. Schilpp, 3-43. La Salle, IL: Open Court.

Chisholm, Roderick. 1976.  Person and Object. La Salle, IL: Open Court.

Cleugh, Mary F. 1937. Time and Its Importance in Modern Thought. London: Methuen.

Davies, P. C. W. 1976. The Physics of Time Asymmetry. Berkeley: University of California Press.

De Beauregard, Costa. 1966. “Time in Relativity Theory: Arguments for a Philosophy of Being.” In Fraser (1966, 417-33).

De Broglie, Louis. 1949. “A General Survey of the Scientific Work of Albert Einstdn.” In Albert Einstein: Philosopher-Scientist, ed. P. A. Schilpp, 107-27. La Salle, IL: Open Court.

Denbigh, Kenneth G. 1975. An Inventive Universe. New York: Braziller.

_______. 1981. Three Concepts of Time. New York: Springer-Verlag.

Dossey, Larry, M.D. 1982. Space, Time and Medicine. With a Foreword by Fritjof Capra. Boulder: Shambala.

Ebon, Martin. 1968. Prophecy in Our Time. New York: New American Library.

Eddington, Arthur. 1922. The Theory of Relativity and Its Influence on Scientific Thought. Oxford: Clarendon.

_______. 1968. The Nature of the Physical World. Ann Arbor: University of Michigan Press.

Einstein, Albert. 1950. The Meaning of Relativity. 3d ed. Princeton: Princeton University Press.

_______. 1954a. Ideas and Opinions. Ed. Carl Seelig. New York: Crown.

_______. 1954b. Relativity: The Special and the General Theory, Trans. R. W. Lawson. London: Methuen.

_______. 1976. “Comment on Meyerson’s ‘La Deduction Relativiste.’” In Čapek (1976, 363-67).

Eisenbud, Jule. 1956. “Psi and the Problem of the Disconnections in Science.” Journal of the American Society for Psychical Research 50:3-26.

Feigl, Herbert. 1960. “Mind-Body, Not a Pseudoproblem.” In Dimensions of Mind, ed. Sydney Hook, 24-36. New York: New York University Press.

Ferré, Frederick. 1972. “Grünbaum on Temporal Becoming: A Critique.” International Philosophical Quarterly 12: 426-45.

_______. 1976. Shaping the Future: Resources for the Post-Modern World. New York: Harper & Row.

Feynman, Richard. 1965. The Character of Physical Law. Cambridge: M.I.T. Press.

Frank, Phillip. 1976. “Is the Future Already Here?” In Capek (1976, 387-95).

Fraser, J. T. 1966. The Voices of Time. New York: Braziller.

_______. Of Time, Passion and Knowledge. New York: Braziller.

Freeman, Eugene, and Sellars, Wilfrid, eds. 1971. Basic Issues in the Philosophy of Time. La Salle, IL: Open Court.

Gale, Richard M. 1968. The Language of Time. New York: Humanities Press.

Gold, Thomas, ed. 1967. The Nature of Time. Ithaca, NY: Cornell University Press.

Gold, Thomas. 1977. “Relativity and Time.” In The Encyclopedia of Ignorance, ed. R. Duncan and M. Weston Smith. New York: Pergamon.

Griffin, David Ray. 1976. God, Power, and Evil: A Process Theodicy. Philadelphia: Westminster.

_______. 1981. “Creation Out of Chaos and the Problem of Evil.” In Encountering Evil: Live Options in Theodicy, ed. Stephen T. Davis, 101-19.  Atlanta: John Knox.

Grünbaum, Adolf. 1963. Philosophical Problems of Space and Time. New York: Knopf.

_______. 1967a. “The Anisotropy of Time.” In Gold (1967, 149-86).

_______. 1967b. Modern Science and Zeno’s Paradoxes. Middletown: Wesleyan University Press.

Hoffman, Banesh (with Helen Dukas). 1972. Albert Einstein: Creator and Rebel. New York: Viking Press.

Lawrence, Nathaniel. 1971. “Time Represented as Space.” In Freeman and Sellars (1971, 123-32).

Lewis, G. N. 1926. “The Nature of Light.” Proceedings of the National Academy of Sciences 12:22-29.

_______. 1930. “The Symmetry of Time in Physics.” Science 71:569-77.

Matson, Floyd W. 1966. The Broken Image: Man, Science and Society. Garden City: Doubleday.

Mehlberg, Henry. 1980. Time, Causality, and the Quantum Theory, vol. I. Ed. Robert S. Cohen. Dordrecht, Holland: Reidel.

Merrill, A. A. 1922. “The t of Physics,” Journal of Philosophy 19 (9), April 27, 238-41.

Meyerson, Emile. 1930. Identity and Reality. Trans. Kate Loewenberg. London: Allen & Unwin.

_______. 1976. “Various Interpretations of Relativistic Time.” In Capek (1976, 353-62).

Monod, Jacques. 1972. Chance and Necessity. New York: Vintage Books.

Needham, Joseph. 1943. Time: The Refreshing River. New York: Macmillan.

Ogden, Schubert. 1966. “Theology and Objectivity.” In The Reality of God and Other Essays, 71-98. New York: Harper & Row.

Popper, Karl R. 1977. (with John C. Eccles), The Self and Its Brain (Springer).

_______. 1982. Postscript to the Logic of Scientific Discovery, ed. W. W. Bartley, III (Rowman and Littlefield), Vol. II: The Open Universe: An Argument for Indeterminism; Vol. III: Quantum Theory and the Schism in Physics.

Prigogine, Ilya. 1973. “Time, Irreversibility and Structure.” In The Physicist’s Conception of Nature, ed. Jagdish Mehra, 561-93. Dordrecht, Holland: Reidel.

_______. 1980. From Being to Becoming. San Francisco: W. H. Freeman.

Prigogine, lIya, and Stengers, Isabelle. 1984. Order Out of Chaos: Man’s New Dialogue with Nature. New York: Bantam Books.

Quine, Willard. 1960. Word and Object. Cambridge: M.I.T. Press.

Reichenbach, Hans. 1956. The Direction of Time. Berkeley: University of California Press.

_______. 1958. The Philosophy of Space and Time. New York: Dover.

Russell, Bertrand. 1917. Mysticism and Logic. London: Allen & Unwin.

_______. 1921. Our Knowledge of the External World. London: Allen and Unwin.

Sipfle, David A. 1971. “On the Intelligibility of the Epochal Theory of Time.” In Freeman and Sellars (1971, 181-94).

Sperry, Roger W. 1976. “Mental Phenomena as Causal Determinants in Brain Function.” In Consciousness and the Brain: A Scientific and Philosophical Inquiry, ed. G. Globus, G. Maxwell, and I. Savodnik, 163-77. New York: Plenum Press.

Steuwer, Roger. 1975. “G. N. Lewis on Detailed Balancing, the Symmetry of Time, and the Nature of Light.” In Historical Studies in the Physical Sciences, vol. 6, ed. Russell McCormmach, 469-511. Princeton: Princeton University Press.

Suppes, Patrick, ed. 1973. Space, Time and Geometry. Dordrecht, Holland: Reidel.

Tanagras, Angelos. 1967. Psychophysical Elements in Parapsychological Traditions. Parapsychology Foundation.

Toulmin, Stephen. 1982. The Return to Cosmology: Postmodern Science and the Theology of Nature. Berkeley: University of California Press.

Toulmin, Stephen, and Goodfield, June. 1965. The Discovery of Time. New York: Harper & Row.

Weyl, Herman. 1949. Philosophy of Mathematics and Natural Science. Princeton: Princeton University Press.

Wheeler, John Archibald. 1982. “Bohr, Einstein, and the Strange Lesson of the Quantum.” In Mind in Nature: Nobel Conference XVII, ed. Richard Q. Elvee, 1-30. New York: Harper & Row.

Whitehead, Alfred North. 1926. Science and the Modern World. 2d ed. New York: Macmillan.

_______. 1927. “Time.” Proceedings of the Sixth International Congress of Philosophy, 59-64. New York and London: Longmans, Green.

_______. 1933. Adventures of Ideas. New York: Macmillan.

_______. 1947. Essays in Science and Philosophy. New York: Philosophical Library.

_______. 1958. The Function of Reason. Boston: Beacon Press.

_______. 1959. Symbolism: Its Meaning and Effect. Capricorn.

_______. 1964. The Concept of Nature. Cambridge: Cambridge University Press.

_______. 1966. Modes of Thought. New York: Free Press.

_______. 1978. Process and Reality. Corrected edition, ed. David Ray Griffin and Donald W. Sherburne. New York: Free Press.

Whitrow, G. J. 1980. The Natural Philosophy of Time. 2d ed. Oxford: Clarendon.

Wilber, Ken, ed. 1982. The Holographic Paradigm and Other Paradoxes. Boulder: Shambala.

Wolf, Fred. 1981. Taking the Quantum Leap: The New Physics for Non-Scientists. San Francisco: Harper & Row.

_______. 1984. Star Wave: Mind, Consciousness, and Quantum Physics. New York: Macmillan.

Zukav, Gary. 1979. The Dancing Wu Li Masters: An Overview of the New Physics. London: Rider/Hutchinson.

Zwart, P. J. 1973. “The Flow of Time.” In Suppes (1973, 131-56).

_______. 1976. About Time, A Philosophical Inquiry into the Origin and Nature of Time. Amsterdam: North-Holland.

Posted May 8, 2007


David Ray Griffin Page