The Anderson Institute Logo 
   Where history is becoming an experimental science      
Innovation and Excellence in Time Technology The Official Site of the World Encyclopedia of Time, the Time Shop, and the Time Research Association.    
   
  Home  |  About Us  |  Educational Resources  |  Encyclopedia  |  The Time Shop  |  Time Research Association  |  Contact Us

Contemporary Philosophies of Time

Time in 20th-century philosophy of physics

Time in the special theory of relativity

Since the classic interpretation of Einstein's special theory of relativity by Hermann Minkowski, a Lithuanian-German mathematician, it has been clear that physics has to do not with two entities, space and time, taken separately, but with a unitary entity space–time, in which, however, timelike and spacelike directions can be distinguished. The Lorentz transformations, which in special relativity define shifts in velocity perspectives, were shown by Minkowski to be simply rotations of space–time axes. The Lorentz contraction of moving rods and the time dilatation of moving clocks turns out to be analogous to the fact that different-sized slices of a sausage are obtained by altering the direction of the slice: just as there is still the objective (absolute) sausage, so also Minkowski restores the absolute to relativity in the form of the invariant four-dimensional object, and the invariance (under the Lorentz transformation) of the space–time interval and of certain fundamental physical quantities such as action (which has the dimensions of energy times time, even though neither energy nor time is separately invariant).

Process philosophers charge the Minkowski universe with being a static one. The philosopher of the manifold denies this charge, saying that a static universe would be one in which all temporal cross sections were exactly similar to one another and in which all particles (considered as four-dimensional objects) lay along parallel lines. The actual universe is not like this, and that it is not static is shown in the Minkowski picture by the dissimilarity of temporal cross sections and the nonparallelism of the world lines of particles. The process philosopher may say that change, as thus portrayed in the Minkowski picture (e.g., with the world lines of particles at varying distances from one another), is not true Bergsonian change, so that something has been left out. But if time advances up the manifold, this would seem to be an advance with respect to a hypertime, perhaps a new time direction orthogonal to the old one. Perhaps it could be a fifth dimension, as has been used in describing the de Sitter universe as a four-dimensional hypersurface in a five-dimensional space. The question may be asked, however, what advantage such a hypertime could have for the process philosopher and whether there is process through hypertime. If there is, one would seem to need a hyper-hypertime, and so on to infinity. (The infinity of hypertimes was indeed postulated by John William Dunne, a British inventor and philosopher, but the remedy seems to be a desperate one.) And if no such regress into hypertimes is postulated, it may be asked whether the process philosopher would not find the five-dimensional universe as static as the four-dimensional one. The process philosopher may therefore adopt the expedient of Henri Bergson, saying that temporal process (the extra something that makes the difference between a static and a dynamic universe) just cannot be pictured spatially (whether one supposes four, five, or more dimensions). According to Bergson, it is something that just has to be intuited and cannot be grasped by discursive reason. The philosopher of the manifold will find this unintelligible and will in any case deny that anything dynamic has been left out of his world picture. This sort of impasse between process philosophers and philosophers of the manifold seems to be characteristic of the present-day state of philosophy.

The theory of relativity implies that simultaneity is relative to a frame of axes. If one frame of axes is moving relative to another, then events that are simultaneous relative to the first are not simultaneous relative to the second, and vice versa. This paradox leads to another difficulty for process philosophy over and above those noted earlier. Those who think that there is a continual coming into existence of events (as the present rushes onward into the future) can be asked “Which present?” It therefore seems difficult to make a distinction between a real present (and perhaps past) as against an as-yet-unreal future. Philosophers of the manifold also urge that to talk of events becoming (coming into existence) is not easily intelligible. Enduring things and processes, in this view, can come into existence; but this simply means that as four-dimensional solids they have an earliest temporal cross section or time slice.

When talking in the fashion of Minkowski, it is advisable, according to philosophers of the manifold, to use tenseless verbs (such as the “equals” in “2 + 2 equals 4”). One can say that all parts of the four-dimensional world exist (in this tenseless sense). This is not, therefore, to say that they all exist now, nor does it mean that Minkowski events are “timeless.” The tenseless verb merely refrains from dating events in relation to its own utterance.

The power of the Minkowski representation is illustrated by its manner in dealing with the so-called clock paradox, which deals with two twins, Peter and Paul. Peter remains on Earth (regarded as at rest in an inertial system) while Paul is shot off in a rocket at half the velocity of light, rapidly decelerated at Alpha Centauri (about four light-years away), and shot back to Earth again at the same speed. Assuming that the period of turnabout is negligible compared with those of uniform velocity, Paul, as a four-dimensional object, lies along the sides AC and CB of a space–time triangle, in which A and B are the points of his departure and return and C that of his turnaround. Peter, as a four-dimensional object, lies along AB. Now, special relativity implies that on his return Paul will be rather more than two years younger than Peter. This is a matter of two sides of a triangle not being equal to the third side: AC + CB < AB. The “less than”—symbolized < —arises from the semi-Euclidean character of Minkowski space–time, which calls for minus signs in its metric (or expression for the interval between two events, which is ds = √(c2dt2 - dx2 - dy2 - dz2) ). The paradox has been held to result from the fact that, from Paul's point of view, it is Peter who has gone off and returned; and so the situation is symmetrical, and Peter and Paul should each be younger than the other—which is impossible. This is to forget, however, the asymmetry reflected in the fact that Peter has been in only one inertial system throughout, and Paul has not; Paul lies along a bent line, Peter along a straight one.

Time in general relativity and cosmology

In general relativity, which, though less firmly established than the special theory, is intended to explain gravitational phenomena, a more complicated metric of variable curvature is employed, which approximates to the Minkowski metric in empty space far from material bodies. Cosmologists who have based their theories on general relativity have sometimes postulated a finite but unbounded space–time (analogous, in four dimensions, to the surface of a sphere) as far as spacelike directions are concerned, but practically all cosmologists have assumed that space–time is infinite in its timelike directions. Kurt Gödel, a contemporary mathematical logician, however, has proposed solutions to the equations of general relativity whereby timelike world lines can bend back on themselves. Unless one accepts a process philosophy and thinks of the flow of time as going around and around such closed timelike world lines, it is not necessary to think that Gödel's idea implies eternal recurrence. Events can be arranged in a circle and still occur only once.

The general theory of relativity predicts a time dilatation in a gravitational field, so that, relative to someone outside of the field, clocks (or atomic processes) go slowly. This retardation is a consequence of the curvature of space–time with which the theory identifies the gravitational field. As a very rough analogy, a road may be considered that, after crossing a plain, goes over a mountain. Clearly, one mile as measured on the humpbacked surface of the mountain is less than one mile as measured horizontally. Similarly—if “less” is replaced by “more” because of the negative signs in the expression for the metric of space–time—one second as measured in the curved region of space–time is more than one second as measured in a flat region. Strange things can happen if the gravitational field is very intense. It has been deduced that so-called black holes in space may occur in places where extraordinarily massive or dense aggregates of matter exist, as in the gravitational collapse of a star. Nothing, not even radiation, can emerge from such a black hole. A critical point is the so-called Schwarzschild radius measured outward from the centre of the collapsed star—a distance, perhaps, of the order of 10 kilometers. Something falling into the hole would take an infinite time to reach this critical radius, according to the space–time frame of reference of a distant observer, but only a finite time in the frame of reference of the falling body itself. From the outside standpoint the fall has become frozen. But from the point of view of the frame of the falling object, the fall continues to zero radius in a very short time indeed—of the order of only 10 or 100 microseconds. Within the black hole spacelike and timelike directions change over, so that to escape again from the black hole is impossible for reasons analogous to those that, in ordinary space–time, make it impossible to travel faster than light. (To travel faster than light a body would have to lie—as a four-dimensional object—in a spacelike direction instead of a timelike one.)

As a rough analogy two country roads may be considered, both of which go at first in a northerly direction. But road A bends round asymptotically toward the east; i.e., it approaches ever closer to a line of latitude. Soon road B crosses this latitude and is thus to the north of all parts of road A. Disregarding the Earth's curvature, it takes infinite space for road A to get as far north as that latitude on road B; i.e., near that latitude an infinite number of “road A northerly units” (say, miles) correspond to a finite number of road B units. Soon road B gets “beyond infinity” in road A units, though it need be only a finite road.

Rather similarly, if a body should fall into a black hole, it would fall for only a finite time, even though it were “beyond infinite” time by external standards. This analogy does not do justice, however, to the real situation in the black hole—the fact that the curvature becomes infinite as the star collapses toward a point. It should, however, help to alleviate the mystery of how a finite time in one reference frame can go “beyond infinity” in another frame.

Most cosmological theories imply that the universe is expanding, with the galaxies receding from one another (as is made plausible by observations of the red shifts of their spectra), and that the universe as it is known originated in a primeval explosion at a date of the order of 15 × 109 years ago. Though this date is often loosely called “the creation of the universe,” there is no reason to deny that the universe (in the philosophical sense of “everything that there is”) existed at an earlier time, even though it may be impossible to know anything of what happened then. (There have been cosmologies, however, that suggest an oscillating universe, with explosion, expansion, contraction, explosion, etc., ad infinitum.) And a fortiori, there is no need to say—as Augustine did in his Confessions as early as the 5th century AD—that time itself was created along with the creation of the universe, though it should not too hastily be assumed that this would lead to absurdity, because common sense could well be misleading at this point.

A British cosmologist, E.A. Milne, however, proposed a theory according to which time in a sense could not extend backward beyond the creation time. According to him there are two scales of time, “τ time” and “t time.” The former is a time scale within which the laws of mechanics and gravitation are invariant, and the latter is a scale within which those of electromagnetic and atomic phenomena are invariant. According to Milne τ is proportional to the logarithm of t (taking the zero of t to be the creation time); thus, by τ time the creation is infinitely far in the past. The logarithmic relationship implies that the constant of gravitation G would increase throughout cosmic history. (This increase might have been expected to show up in certain geological data, but apparently the evidence is against it.)

Time in microphysics

Special problems arise in considering time in quantum mechanics and in particle interactions.
Quantum-mechanical aspects of time
In quantum mechanics it is usual to represent measurable quantities by operators in an abstract many-dimensional (often infinite-dimensional) so-called Hilbert space. Nevertheless, this space is an abstract mathematical tool for calculating the evolution in time of the energy levels of systems—and this evolution occurs in ordinary space–time. For example, in the formula AH - HA = iℏ(dA/dt), in which i is √(−1) and ℏ is 1/2π times Planck's constant, h, the A and H are operators, but the t is a perfectly ordinary time variable. There may be something unusual, however, about the concept of the time at which quantum-mechanical events occur, because according to the Copenhagen interpretation of quantum mechanics the state of a microsystem is relative to an experimental arrangement. Thus energy and time are conjugate: no experimental arrangement can determine both simultaneously, for the energy is relative to one experimental arrangement, and the time is relative to another. (Thus, a more relational sense of “time” is suggested.) The states of the experimental arrangement cannot be merely relative to other experimental arrangements, on pain of infinite regress; and so these have to be described by classical physics. (This parasitism on classical physics is a possible weakness in quantum mechanics over which there is much controversy.)

The relation between time uncertainty and energy uncertainty, in which their product is equal to or greater than h/4π, ΔEΔt ⋜ h/4π, has led to estimates of the theoretical minimum measurable span of time, which comes to something of the order of 10-24 second and hence to speculations that time may be made up of discrete intervals (chronons). These suggestions are open to a very serious objection, viz., that the mathematics of quantum mechanics makes use of continuous space and time (for example, it contains differential equations). It is not easy to see how it could possibly be recast so as to postulate only a discrete space–time (or even a merely dense one). For a set of instants to be dense, there must be an instant between any two instants. For it to be a continuum, however, something more is required, viz., that every set of instants earlier (later) than any given one should have an upper (lower) bound. It is continuity that enables modern mathematics to surmount the paradox of extension framed by the Pre-Socratic Eleatic Zeno—a paradox comprising the question of how a finite interval can be made up of dimensionless points or instants.
Time in particle interactions
Until recently it was thought that the fundamental laws of nature are time symmetrical. It is true that the second law of thermodynamics, according to which randomness always increases, is time asymmetrical; but this law is not strictly true (for example, the phenomenon of Brownian motion contravenes it), and it is now regarded as a statistical derivative of the fundamental laws together with certain boundary conditions. The fundamental laws of physics were long thought also to be charge symmetrical (for example, an antiproton together with a positron behave like a proton and electron) and to be symmetrical with respect to parity (reflection in space, as in a mirror). The experimental evidence now suggests that all three symmetries are not quite exact but that the laws of nature are symmetrical if all three reflections are combined: charge, parity, and time reflections forming what can be called (after the initials of the three parameters) a CPT mirror. The time asymmetry was shown in certain abstruse experiments concerning the decay of K mesons that have a short time decay into two pions and a long time decay into three pions.

Time in molar physics

The above-mentioned violations of temporal symmetry in the fundamental laws of nature are such out-of-the-way ones, however, that it seems unlikely that they are responsible for the gross violations of temporal symmetry that are apparent in the visible world. An obvious asymmetry is that there are traces of the past (footprints, fossils, tape recordings, memories) and not of the future. There are mixing processes but no comparable unmixing process: milk and tea easily combine to give a whitish brown liquid, but it requires ingenuity and energy and complicated apparatus to separate the two liquids. A cold saucepan of water on a hot brick will soon become a tepid saucepan on a tepid brick; but the heat energy of the tepid saucepan never goes into the tepid brick to produce a cold saucepan and a hot brick. Even though the laws of nature are assumed to be time symmetrical, it is possible to explain these asymmetries by means of suitable assumptions about boundary conditions. Much discussion of this problem has stemmed from the work of Ludwig Boltzmann, an Austrian physicist, who showed that the concept of the thermodynamic quantity entropy could be reduced to that of randomness or disorder. Among 20th-century philosophers in this tradition may be mentioned Hans Reichenbach, a German-U.S. Positivist, Adolf Grünbaum, a U.S. philosopher, and Olivier Costa de Beauregard, a French philosopher-physicist. There have also been many relevant papers of high mathematical sophistication scattered through the literature of mathematical physics. Reichenbach (and Grünbaum, who improved on Reichenbach in some respects) explained a trace as being a branch system; i.e., a relatively isolated system, the entropy of which is less than would be expected if one compared it with that of the surrounding region. For example, a footprint on the beach has sand particles compressed together below a volume containing air only, instead of being quite evenly (randomly) spread over the volume occupied by the compressed and empty parts.

Another striking temporal asymmetry on the macro level, viz., that spherical waves are often observed being emitted from a source but never contracting to a sink, has been stressed by Sir Karl Popper, a 20th-century Austrian and British philosopher of science. By considering radiation as having a particle aspect (i.e., as consisting of photons), Costa de Beauregard has argued that this “principle of retarded waves” can be reduced to the statistical Boltzmann principle of increasing entropy and so is not really different from the previously discussed asymmetry. These considerations also provide some justification for the common-sense idea that the cause–effect relation is a temporally unidirectional one, even though the laws of nature themselves allow for retrodiction no less than for prediction.

A third striking asymmetry on the macro level is that of the apparent mutual recession of the galaxies, which can plausibly be deduced from the red shifts observed in their spectra. It is still not clear whether or how far this asymmetry can be reduced to the two asymmetries already discussed, though interesting suggestions have been made.

The statistical considerations that explain temporal asymmetry apply only to large assemblages of particles. Hence, any device that records time intervals will have to be macroscopic and to make use somewhere of statistically irreversible processes. Even if one were to count the swings of a frictionless pendulum, this counting would require memory traces in the brain, which would function as a temporally irreversible recording device.

Time in 20th-century philosophy of biology and philosophy of mind

Organisms often have some sort of internal clock that regulates their behavior. There is a tendency, for example, for leaves of leguminous plants to alter their position so that they lie in one position by day and in another position by night. This tendency persists if the plant is in artificial light that is kept constant, though it can be modified to other periodicities (e.g., to a six-hour instead of a 24-hour rhythm) by suitably regulating the periods of artificial light and darkness. In animals, similar daily rhythms are usually acquired, but in experimental conditions animals nevertheless tend to adapt better to a 24-hour rhythm than to any other. Sea anemones expand and contract to the rhythm of the tides, and this periodic behavior will persist for some time even when the sea anemone is placed in a tank. Bees can be trained to come for food at fixed periods (e.g., every 21 hours), and this demonstrates that they possess some sort of internal clock. Similarly, humans themselves have some power to estimate time in the absence of clocks and other sensory cues. This fact refutes the contention of the 17th-century English philosopher John Locke (and of other philosophers in the Empiricist tradition) that time is perceived only as a relation between successive sensations. The U.S. mathematician Norbert Wiener has speculated on the possibility that the human time sense depends on the α-rhythm of electrical oscillation in the brain.

Temporal rhythms in both plants and animals (including humans) are dependent on temperature, and experiments on human subjects have shown that, if their temperature is raised, they underestimate the time between events.

Despite these facts, the Lockean notion that the estimation of time depends on the succession of sensations is still to some degree true. People who take the drugs hashish and mescaline, for example, may feel their sensations following one another much more rapidly. Because there are so many more sensations than normal in a given interval of time, time seems to drag, so that a minute may feel like an hour. Similar illusions about the spans of time occur in dreams.

It is unclear whether most discussions of so-called biological and psychological time have much significance for metaphysics. As far as the distorted experiences of time that arise through drugs (and in schizophrenia) are concerned, it can be argued that there is nothing surprising in the fact that pathological states can make people misestimate periods of time, and so it can be claimed that facts of this sort do not shed any more light on the philosophy of time than facts about mountains looking near after rainstorms and looking far after dust storms shed on the philosophy of space.

The idea that psychological studies of temporal experience are philosophically important is probably connected with the sort of Empiricism that was characteristic of Locke and still more of the Empiricists George Berkeley and David Hume and their successors. The idea of time had somehow to be constructed out of the primitive experience of ideas succeeding one another. Nowadays, concept formation is thought of as more of a social phenomenon involved in the “picking up” of a language; thus, contemporary philosophers have tended to see the problem differently: humans do not have to construct their concepts from their own immediate sensations. Even so, the learning of temporal concepts surely does at least involve an immediate apprehension of the relation of “earlier” and “later.” A mere succession of sensations, however, will go no way toward yielding the idea of time: if one sensation has vanished entirely before the other is in consciousness, one cannot be immediately aware of the succession of sensations. What Empiricism needs, therefore, as a basis for constructing the idea of time is an experience of succession as opposed to a succession of experiences. Hence, two or more ideas that are related by “earlier than” must be experienced in one single act of awareness. William James, a U.S. Pragmatist philosopher and also a pioneer psychologist, popularized the term specious present for the span of time covered by a single act of awareness. His idea was that at a given moment of time a person is aware of events a short time before that time. (Sometimes he spoke of the specious present as a saddleback looking slightly into the future as well as slightly into the past, but this was inconsistent with his idea that the specious present depended on lingering short-term memory processes in the brain.) He referred to experiments by the German psychologist Wilhelm Wundt that showed that the longest group of arbitrary sounds that a person could identify without error lasted about six seconds. Other criteria perhaps involving other sense modalities might lead to slightly different spans of time, but the interesting point is that, if there is such a specious present, it cannot be explained solely by ordinary memory traces: if one hears a “tick tock” of a clock, the “tick” is not remembered in the way in which a “tick tock” 10 minutes ago is remembered. The specious present is perhaps not really specious: the idea that it was specious depended on an idea that the real (nonspecious) present had to be instantaneous. If perception is considered as a certain reliable way of being caused to have true beliefs about the environment by sensory stimulation, there is no need to suppose that these true beliefs have to be about an instantaneous state of the world. It can therefore be questioned whether the term specious is a happy one.

Two matters discussed earlier in connection with the philosophy of physics have implications for the philosophy of mind: (1) the integration of space and time in the theory of relativity makes it harder to conceive of immaterial minds that exist in time but are not even localizable in space; (2) the statistical explanation of temporal asymmetry explains why the brain has memory traces of the past but not of the future and, hence, helps to explain the unidirectional nature of temporal consciousness. It also gives reasons for skepticism about the claims of parapsychologists to have experimental evidence for precognition; or it shows, at least, that if these phenomena do exist they are not able to be fitted into a cosmology based on physics as it exists today.