You are on page 1of 83

Time

Time has been studied by philosophers and scientists for 2,500 years, and thanks to
this attention it is much better understood today. Nevertheless, many issues remain to
be resolved. Here is a short list of the most important ones—what time actually is;
whether time exists when nothing is changing; what kinds of time travel are possible;
why time has an arrow; whether the future and past are real; how to analyze the
metaphor of time’s flow; whether future time will be infinite; whether there was time
before the Big Bang; whether tensed or tenseless concepts are semantically basic;
what is the proper formalism or logic for capturing the special role that time plays in
reasoning; and what are the neural mechanisms that account for our experience of
time. Some of these issues will be resolved by scientific advances alone, but others
require philosophical analysis.

Consider this one issue upon which philosophers of time are deeply divided: What sort
of ontological differences are there among the present, past and future? There are
three competing theories. Presentists argue that necessarily only present objects and
present experiences are real, and we conscious beings recognize this in the special
“vividness” of our present experience. The dinosaurs have slipped out of reality.
According to the growing-universe or growing-block theory, the past and present are
both real, but the future is not because the future is indeterminate or merely potential.
Dinosaurs are real, but our death is not. The third and more popular theory is that there
are no significant ontological differences among present, past, and future because the
differences are merely subjective. This view is called “the block universe theory” or
“eternalism.”

That controversy raises the issue of tenseless versus tensed theories of time. The block
universe theory implies a tenseless theory. The earliest version of this theory implied
that tensed terminology can be removed and replaced with tenseless terminology. For
example, the future-tensed sentence, “The Lakers will win the basketball game” might
be analyzed as, “The Lakers do win at time t, and time t happens after the time of this
utterance.” The future tense has been removed, and the new verb phrases “do win”
and “happens after” are tenseless logically, although they are grammatically in the
present tense. Advocates of a tensed theory of time object to this strategy and say that
tenseless terminology is not semantically basic but should be analyzed in tensed terms,
and that tensed facts are needed to make the tensed statements be true. For example,
a tensed theory might imply that no adequate account of the present tensed fact that it
is now midnight can be given without irreducible tensed properties such as presentness
or now-ness. So, the philosophical debate is over whether tensed concepts have
semantical priority over untensed concepts, and whether tensed facts have ontological
priority over untensed facts.

This article explores both what is now known about time and what is controversial and
unresolved, by addressing the questions listed in the table of contents.

Table of Contents

1. What Should a Philosophical Theory of Time Do?


2. How is Time Related to Mind?
3. What is Time?
1. The Variety of Answers
2. Time vs. “Time”
3. Defining Time Order with Causal Order
4. Linear and Circular Time
5. Does Time Emerge from Something More Basic?
4. What does Science Require of Time?
1. Relativity and Quantum Mechanics
2. The Big Bang
3. Infinite Time
4. Atoms of Time
5. What Kinds of Time Travel are Possible?
6. Is the Relational Theory Preferable to the Absolute Theory?
7. Does Time Flow?
8. What Gives Time its Direction or “Arrow”?
1. What Needs to be Explained?
2. Explanations or Theories of the Arrow
3. Multiple Arrows
4. Reversing Time
9. Is Only the Present Real?
10. Are there Essentially Tensed Facts?
11. What is Temporal Logic?
12. Supplements
1. Frequently Asked Questions
2. Special Relativity: Proper Times, Coordinate Systems, and Lorentz
Transformations
13. References and Further Reading

1. What Should a Philosophical Theory of Time Do?


Should it define the word “time”? Yes, but it is improper to demand that we define our
term “time” as a prelude to saying anything more about time, in large part because, as
we have learned more about time, our definition has evolved. What we really want is to
build a comprehensive, philosophical theory of time that helps us understand time by
solving problems about time. We do not want to start building this theory by adopting a
definition of time that prejudices the project from the beginning.

Although there are theories of how to solve a specific problem about time, it is always
better to knit together solutions to several problems. Ideally, the goal is to produce a
theory of time that will solve in a systematic way the constellation of problems
involving time. What are those problems?

One is to clarify the relationship between time and the mind. Does time exist for beings
that have no minds? It is easy to confuse time itself with the perception of time.

Another problem is to decide which of our intuitions about time should be retained.
Some of these intuitions may reflect deep insights into the nature of time, and others
may be faulty ideas inherited from our predecessors. It is not obvious which is which.
For one example, if we have the intuition that time flows, but our science implies
otherwise, then which view should get priority? Philosophers of time must solve the
problem of how to treat our temporal intuitions.
A third problem for a philosophical theory of time is to clarify what physical science
presupposes and implies about time. A later section of this article examines this topic.
Most all philosophers of time claim that philosophical theories should be consistent with
physical science, or, if not, then they must accept the heavy burden of proof to justify
the inconsistency.

A philosophical theory of time should describe the relationship between instants and
events. Does the instant that we label as “11:01 A.M.” for a certain date exist
independently of the events that occur then? In other words, can time exist if no event
is happening? This question or problem raises the thorny metaphysical issue of
absolute vs. relational theories of time.

A theory of time should address the question of time’s apparent direction. If the
projectionist in the movie theater (cinema) shows a film of cream being added into
black coffee but runs the film backwards, we in the audience can immediately tell that
events could not have occurred this way. We recognize the arrow of time because we
know about the one-directional processes in nature. This arrow or unidirectionality
becomes less and less apparent to us viewers as the film subject gets smaller and
smaller and the time interval gets shorter and shorter until finally we are viewing
processes that could just as easily go the other way, at which point the arrow of time
has disappeared. Philosophers disagree about the explanation of the arrow. Could it be
a consequence of the laws of science? The arrow appears to be very basic for
understanding nature, yet it is odd that asymmetries in time do not appear in the
principal, basic dynamical laws of physics. Could the arrow of time reverse some day?
Philosophers wonder what life would be like in some far off corner of the universe if the
arrow of time were reversed there. Would people there walk backwards up steps while
remembering the future?

Another philosophical problem about time concerns the two questions, “What is the
present, and why does it move into the past?” If we know what the present is, then we
ought to be able to answer the question, “How long does the present last?” Regarding
the “movement” of the present into the past, many philosophers are suspicious of this
notion of the flow of time, the march of time. They doubt whether it is a property of
time as opposed to being some feature of human perception. Assuming time does flow,
is the flow regular? If the flow is irregular, then perhaps Friday seconds last longer than
Thursday seconds, as the flow of Friday time slows to a crawl, or perhaps Friday might
contain more seconds than Thursday.

Are there ontological differences among the past, present, and future? Some
philosophers doubt whether the future and past are as real as the present, the feature
that is referred to by the word “now.” A famous philosophical argument says that, if the
future were real, then it would be fixed now, and we would not have the freedom to
affect that future. Since we do have that freedom, the future can not be real. Some
philosophers consider this to be a clever, but faulty argument.

For a last example of a philosophical issue regarding time, is time a fundamental


feature of nature, or does it emerge from more basic features–in analogy to the way
the smoothness of water flow emerges from the complicated behavior of the underlying
molecules? From what more basic feature does time emerge?
A full theory of time should address this constellation of philosophical issues about
time. Narrower theories of time will focus on resolving a few members of this
constellation, but the long-range goal is to knit together these theories into a full,
systematic, detailed theory of time.

2. How is Time Related to Mind?


Physical time is public time, the time that clocks are designed to measure.
Psychological time or phenomenological time is private time. It is perhaps best
understood as awareness of physical time. Psychological time passes swiftly for us
while we are enjoying reading a book, but it slows dramatically if we are waiting
anxiously for the water to boil on the stove. The slowness is probably due to focusing
our attention on short intervals of physical time. Meanwhile, the clock by the stove is
measuring physical time and is not affected by anybody’s awareness.

When a physicist defines speed to be the rate of change of position with respect to
time, the term “time” refers to physical time. Physical time is more basic for helping us
understand our shared experiences in the world, and so it is more useful than
psychological time for doing science. But psychological time is vitally important for
understanding many human thought processes. We have an awareness of the passage
of time even during our sleep, and we awake knowing we have slept for one night, not
for one month. But if we have been under a general anesthetic or have been knocked
unconscious and then wake up, we may have no sense of how long we have been
unconscious. Psychological time stopped. Some philosophers claim that psychological
time is completely transcended in the mental state called “nirvana.”

Within the field of cognitive science, one wants to know what are the neural
mechanisms that account not only for our experience of time’s flow, but also for our
ability to place events into the proper time order. See (Damasio, 2006) for further
discussion of the progress in this area of cognitive science. The most surprising
scientific discovery about psychological time is Benjamin Libet’s experiments in the
1970s that show, or so it is claimed, that the brain events involved in initiating free
choices occur about a third of a second before we are aware of our choice. Before
Libet’s work, it was universally agreed that a person is aware of deciding to act freely,
then later the body initiates the action.

Psychologists are interested in whether we can speed up our minds relative to physical
time. If so, we might become mentally more productive, get more high quality decision
making done per fixed amount of physical time, learn more per minute. Several
avenues have been explored: using drugs such as cocaine and amphetamines,
undergoing extreme experiences such as jumping backwards off a tall tower with
bungee cords attached to the legs, and trying different forms of meditation. So far,
none of these avenues have led to success productivity-wise.

Any organism’s sense of time is subjective, but is the time that is sensed also
subjective, a mind-dependent phenomenon? Without minds in the world, nothing in the
world would be surprising or beautiful or interesting. Can we add that nothing would be
in time? If judgments of time were subjective in the way judgments of being interesting
vs. not-interesting are subjective, then it would be miraculous that everyone can so
easily agree on the ordering of public events in time. For example, first, Einstein was
born, then he went to school, then he died. Everybody agrees that it happened in this
order: birth, school, death. No other order. The agreement on time order for so many
events is part of the reason that most philosophers and scientists believe physical time
is an objective phenomenon not dependent on being consciously experienced. The
other part of the reason time is believed to be objective is that our universe has a large
number of different processes that bear consistent time relations, or frequency of
occurrence relations, to each other. For example, the frequency of a fixed-length
pendulum is a constant multiple of the half life of a specific radioactive uranium
isotope; the relationship does not change as time goes by (at least not much and not
for a long time). The existence of these sorts of relationships makes our system of
physical laws much simpler than it otherwise would be, and it makes us more confident
that there is something objective we are referring to with the time-variable in those
laws. The stability of these relationships over a long time also make it easy to create
clocks. Time can be measured easily because we have access to long term simple
harmonic oscillators that have a regular period or “regular ticking.” This regular motion
shows up in completely different stable systems when they are disturbed: a ball
swinging from a string (a pendulum), a ball bouncing up and down from a coiled spring,
a planet orbiting the sun, organ pipes, electric circuits, and atoms in a crystal lattice.
Many of these systems make good clocks.

Aristotle raised this issue of the mind-dependence of time when he said, “Whether, if
soul (mind) did not exist, time would exist or not, is a question that may fairly be
asked; for if there cannot be someone to count there cannot be anything that can be
counted…” [Physics, chapter 14]. He does not answer his own question because, he
says rather profoundly, it depends on whether time is the conscious numbering of
movement or instead is just the capability of movements being numbered were
consciousness to exist.

St. Augustine, adopting a subjective view of time, said time is nothing in reality but
exists only in the mind’s apprehension of that reality. In the 11th century, the Persian
philosopher Avicenna doubted the existence of physical time, arguing that time exists
only in the mind due to memory and expectation. The 13th century philosophers Henry
of Ghent and Giles of Rome said time exists in reality as a mind-independent
continuum, but is distinguished into earlier and later parts only by the mind. In the 13th
century, Duns Scotus clearly recognized both physical and psychological time.

At the end of the 18th century, Kant suggested a subtle relationship between time and
mind–that our mind actually structures our perceptions so that we can know a priori
that time is like a mathematical line. Time is, on this theory, a form of conscious
experience, and our sense of time is a necessary condition of our experience. In the
19th century, Ernst Mach claimed instead that our sense of time is a simple sensation.
This controversy took another turn when other philosophers argued that both Kant and
Mach were incorrect because our sense of time is an intellectual construction (see
Whitrow, p. 64).

In the 20th century, the philosopher of science Bas van Fraassen described physical
time by saying, “There would be no time were there no beings capable of reason” just
as “there would be no food were there no organisms, and no teacups if there were no
tea drinkers,” and no cultural objects without a culture.
The controversy in metaphysics between idealism and realism is that, for the idealist,
nothing exists independently of the mind. If this controversy is settled in favor of
idealism, then time, too, would have that subjective feature–physical time as well as
psychological time.

It has been suggested by some philosophers that Einstein’s theory of relativity, when
confirmed, showed us that time depends on the observer, and thus that time is
subjective, or dependent on the mind. This error is probably caused by Einstein’s use of
the term “observer.” Einstein’s theory does imply that the duration of an event is not
absolute but depends on the observer’s frame of reference or coordinate system. But
what Einstein means by “observer’s frame of reference” is merely a perspective or
framework from which measurements could be made. The “observer” does not have to
be a conscious being or have a mind. So, Einstein is not making a point about mind-
dependence.

For more on the consciousness of time and related issues, see the article
“Phenomenology and Time-Consciousness.”

3. What is Time?
a. The Variety of Answers

The most popular short answer to the question “What is physical time?” is that it is not
a substance or object but rather a special system of relations among instantaneous
events. This is the answer offered by Adolf Grünbaum who applies the contemporary
mathematical theory of continuity to physical processes, and says time is a linear
continuum of instants and is a distinguished one-dimensional sub-space of four-
dimensional spacetime.

How do we tell whether this is the correct answer to our question? To be convinced, we
need to be told what the relevant terms mean, such as “certain system of relations.” In
addition, we need to be presented with a theory of time implying that time is this
system of relations; and we need to be shown how that theory adequately addresses
the many features that are required for a successful theory of time. Finally, we need to
compare this theory to its alternatives. This article will not carry out these tasks.

A different, but popular answer to the question “What is time?” is that time is the form
of becoming. To assess this answer, which is from Alfred North Whitehead, we need to
be told what the term “form of becoming” means; we need to be presented with a
detailed theory of time implying that time is the form of becoming; and we need to
investigate how it addresses those many features required for a successful theory of
time. A third answer or theory of time is Michael Dummett’s constructive model of
time; he argues that time is a composition of intervals rather than of durationless
instants. The model is constructive in the sense that it implies there do not exist any
times which are not detectable in principle by a physical process. A fourth answer is
that time is a distinguished one-dimensional sub-space of spacetime, but spacetime is
a substance. This substantivalist answer is explored in a later section. There are many
other ways that our question has been answered.
If physical time and psychological time are two different kinds of time, then two
answers are required to the question “What is time?” and some commentary is
required regarding their relationships, such as whether one is more fundamental. Many
philosophers of science argue that physical time is more fundamental even though
psychological time is discovered first by each of us as we grow out of our childhood,
and even though psychological time was discovered first as we human beings evolved
from our animal ancestors. The remainder of this article focuses more on physical time
than psychological time.

Another answer to our question, “What is time?” is that time is whatever the time
variable t is denoting in the best-confirmed and most fundamental theories of current
science. “Time” is given an implicit definition this way. Nearly all philosophers would
agree that we do learn much about physical time by looking at the behavior of the time
variable in these theories; but they complain that the full nature of physical time can
be revealed only with a philosophical theory of time that addresses the many
philosophical issues that scientists do not concern themselves with.

Bothered by the contradictions they claimed to find in our concept of time, some
philosophers, notably Zeno, Plato, Spinoza, Hegel, and McTaggart, answer the question,
“What is time?” by replying that it is nothing because it does not exist. In a similar vein,
the early 20th century English philosopher F. H. Bradley argues, “Time, like space, has
most evidently proved not to be real, but a contradictory appearance….The problem of
change defies solution.” However, most philosophers agree that time does exist. They
just can not agree on what it is.

Let’s briefly explore other answers that have been given throughout history to our
question, “What is time?” Aristotle claimed that “time is the measure of change”
[Physics, chapter 12], but he emphasized “that time is not change [itself]” because a
change “may be faster or slower, but not time…” [Physics, chapter 10]. For example, a
specific change such as the descent of a leaf can be faster or slower, but time itself can
not be faster or slower. In developing his views about time, Aristotle advocated what is
now referred to as the relational theory when he said, “there is no time apart from
change….” [Physics, chapter 11]. In addition, Aristotle said time is not discrete or
atomistic but “is continuous…. In respect of size there is no minimum; for every line is
divided ad infinitum. Hence it is so with time” [Physics, chapter 11].

René Descartes had a very different answer to “What is time?” He argued that a
material body has the property of spatial extension but no inherent capacity for
temporal endurance, and that God by his continual action sustains (or re-creates) the
body at each successive instant. Time is a kind of sustenance or re-creation.

In the 17th century, the English physicist Isaac Barrow rejected Aristotle’s linkage
between time and change. Barrow said time is something which exists independently of
motion or change and which existed even before God created the matter in the
universe. Barrow’s student, Isaac Newton, agreed that this absolute theory of time is
correct. Newton argued very specifically that time and space are an infinitely large
container for all events, and that the container exists with or without the events. He
added that space and time are not material substances, but are like substances in not
being dependent on matter or motions or anything else except God.
Gottfried Leibniz objected. He argued that time is not an entity existing independently
of actual events. He insisted that Newton had underemphasized the fact that time
necessarily involves an ordering of any pair of non-simultaneous events. This is why
time “needs” events, so to speak. Leibniz added that this overall order is time. He
accepts a relational theory of time and rejects an absolute theory.

In the 18th century, Immanuel Kant said time and space are forms that the mind
projects upon the external things-in-themselves. He spoke of our mind structuring our
perceptions so that space always has a Euclidean geometry, and time has the structure
of the mathematical line. Kant’s idea that time is a form of apprehending phenomena is
probably best taken as suggesting that we have no direct perception of time but only
the ability to experience things and events in time. Some historians distinguish
perceptual space from physical space and say that Kant was right about perceptual
space. It is difficult, though, to get a clear concept of perceptual space. If physical
space and perceptual space are the same thing, then Kant is claiming we know a priori
that physical space is Euclidean. With the discovery of non-Euclidean geometries in the
1820s, and with increased doubt about the reliability of Kant’s method of
transcendental proof, the view that truths about space and time are apriori truths
began to lose favor.

b. Time vs. “Time”

Whatever time is, it is not “time.” One has four letters; the other does not.
Nevertheless, it might help us understand time if we improved our understanding of the
sense of the word “time.” Should the proper answer to the question “What is time?”
produce a definition of the word as a means of capturing its sense? Definitely not–if the
definition must be some analysis that provides a simple paraphrase in all its
occurrences. There are just too many varied occurrences of the word: time out, behind
the times, in the nick of time, and so forth.

But how about narrowing the goal to a definition of the word “time” in its main sense,
the sense that most interests philosophers and physicists? That is, explore the usage of
the word “time” in its principal sense as a means of learning what time is. Well, this
project would require some consideration of the grammar of the word “time.” Most
philosophers today would agree with A. N. Prior who remarked that, “there are genuine
metaphysical problems, but I think you have to talk about grammar at least a little bit
in order to solve most of them.” However, do we learn enough about what time is when
we learn about the grammatical intricacies of the word? John Austin made this point in
“A Plea for Excuses,” when he said, if we are using the analytic method, the method of
analysis of language, in order to sharpen our perception of the phenomena, then “it is
plainly preferable to investigate a field where ordinary language is rich and subtle, as it
is in the pressingly practical matter of Excuses, but certainly is not in the matter, say,
of Time.” Ordinary-language philosophers have studied time talk, what Wittgenstein
called the “language game” of discourse about time. Wittgenstein’s expectation is that
by drawing attention to ordinary ways of speaking we will be able to dissolve rather
than answer our philosophical questions. But most philosophers of time are unsatisfied
with this approach; they want the questions answered, not dissolved, although they are
happy to have help from the ordinary language philosopher in clearing up
misconceptions that may be produced by the way we use the word in our ordinary,
non-technical discourse.
c. Defining Time Order with Causal Order

In 1924, Hans Reichenbach defined time order in terms of possible cause. Event A
happens before event B if A could have caused B but B could not have caused A. This
was the first causal theory of time, although Leibniz had said, “If of two elements which
are not simultaneous one comprehends the cause of the other, then the former is
considered as preceding, the latter as succeeding.” The usefulness of the causal theory
depends on a clarification of the notorious notions of causality and possibility without
producing a circular explanation that presupposes an understanding of time order.
Reichenbach’s idea was that causal order can be explained in terms of the “fork
asymmetry.” The asymmetry is due to the fact that outgoing processes from a common
center tend to be correlated with one another, but incoming processes to a common
center are uncorrelated. [Do you remember ever tossing a rock into a still pond?
There’s a correlation among all sorts of later events such as the rock’s disappearing
under the water, the water surface getting wavy, your hearing a splash sound, the
water surging slightly up the bank at the edge of the pond, and even of the pond being
warmer. Imagine what the initial conditions at the edge and bottom of the pond must
be like to produce correlated, incoming, concentric water waves so that as they reach
the center the rock flies out of the water, leaving the water surface smooth, and sound
waves rush out of your ear and converge on the surface where the splash is
unoccuring, and the pond is left cooler.] Some philosophers argue that temporal
asymmetry, but not temporal priority, can be analyzed in terms of causation [put more
simply, event A's not occuring simultaneously with B can be analyzed in terms of cause
and possible cause, but what can't be analyzed in this manner is A's occuring first].
Even if Reichenbach were correct that temporal priority can be analyzed in terms of
causation, the question remains whether time itself can be analyzed in those terms.

The usefulness of the causal theory also depends on a refutation of David Hume’s view
that causation is simply a matter of constant conjunction [that is, event A's causing
event B is simply B's always occurring if A does]. For Hume, there is nothing
metaphysically deep about causes preceding their effects; it is just a matter of
convention that we use the terms “cause” and “effect” to distinguish the earlier and
later members of a pair of events which are related by constant conjunction.

d. Linear and Circular Time

During history, a variety of answers have been given to the question of whether time is
like a line or, instead, like a circle. The concept of linear time first appeared in the
writings of the Hebrews and the Zoroastrian Iranians. The Roman writer Seneca also
advocated linear time. Plato and most other Greeks and Romans believed time to be
motion and believed cosmic motion was cyclical, but this was not envisioned as
requiring any detailed endless repetition such as the multiple rebirths of Socrates.
However, the Pythagoreans and some Stoic philosophers did adopt this drastic position.

With circular time, you can be assured that after your death you will be reborn. The
future will become the past. If time is like this, then the question arises as to whether
there would be an endless number of times when each state of the world reoccurred, or
whether, accepting Leibniz’s Principle of the Identity of Indiscernibles, each supposedly
repeating state of the world would occur just once because each state would be not be
discernible from the repeated state.
Islamic and Christian theologians adopted the idea that time is linear plus the Jewish-
Zoroastrian idea that the universe was created at a definite moment in the past.
Augustine emphasized that human experience is a one-way journey from Genesis to
Judgment, regardless of any recurring patterns or cycles in nature. In the Medieval
period, Thomas Aquinas agreed. Nevertheless, it was not until 1602 that the concept of
linear time was more clearly formulated–by the English philosopher Francis Bacon. In
1687, Newton advocated linear time when he represented time mathematically by
using a continuous straight line. The concept of linear time was promoted by Barrow,
Leibniz, Locke and Kant. In 19th century Europe, the idea of linear time became
dominant in both science and philosophy. However, in the twentieth century, Gödel and
others discovered solutions to the equations of Einstein’s general theory of relativity
that allowed closed loops of proper time. These causal loops or closed curves in
spacetime allow you to go forward continuously in time until you arrive back into your
past. You will become your younger self in the future. Gödel believed that even though
our universe does not exemplify this solution to Einstein’s equations, the very
possibility shows that time is unreal because, he believed, the concept of time does not
allow loops. It is an open question in the analysis of the concept of time as to whether
the concept does or does not allow loops.

e. Does Time Emerge from Something More Basic?

Is time ontologically basic, or does it depend on something more basic? We might


rephrase this question as whether facts about time supervene on more basic facts.
Facts about sound supervene on, or are a product of, facts about changes in the
molecules of the air, so molecular change is more basic than sound. Thanks to
Minkowski in 1908 we believe spacetime is more basic than time, but is spacetime
itself basic? Some physicists argue that both space and time are the product of some
more basic micro-substrate, although there is no agreed-upon theory of what the
substrate is. Other physicists say space is not basic, but time is. In 2004, after winning
the Nobel Prize in physics, David Gross expressed this viewpoint:

Everyone in string theory is convinced…that spacetime is doomed. But we don’t know


what it’s replaced by. We have an enormous amount of evidence that space is doomed.
We even have examples, mathematically well-defined examples, where space is an
emergent concept…. But in my opinion the tough problem that has not yet been faced
up to at all is, “How do we imagine a dynamical theory of physics in which time is
emergent?” …All the examples we have do not have an emergent time. They have
emergent space but not time. It is very hard for me to imagine a formulation of physics
without time as a primary concept because physics is typically thought of as predicting
the future given the past. We have unitary time evolution. How could we have a theory
of physics where we start with something in which time is never mentioned?

4. What does Science Require of Time?


a. Relativity and Quantum Mechanics

The general theory of relativity and quantum mechanics are the two most fundamental
theories of physics, and the Big Bang theory is the leading theory of cosmology.
According to relativity and quantum mechanics, spacetime is, loosely speaking, a
collection of points called “spacetime locations” where the universe’s physical events
occur. Spacetime is four-dimensional and a continuum, and time is a distinguished,
one-dimensional sub-space of this continuum. Any interval of time–any duration–is a
linear continuum of instants. So, a duration has a point-like structure similar to the
structure of an interval of real numbers; between any two instants there is another
instant, and there are no gaps in the sequence of instants. This is what science requires
time to be, but we haven’t commented on why science requires time to be this way.

This first response to the question “What does science require of time?” is too simple.
There are complications. There is an important difference between the universe’s
cosmic time and a clock’s proper time; and there is an important difference between
proper time and a reference frame’s coordinate time. Most spacetimes can not have
coordinate systems. Also, all physicists believe that relativity and quantum mechanics
are logically inconsistent and need to be replaced by a theory of quantum gravity. A
theory of quantum gravity is likely to have radical implications for our understanding of
time, such as time and space being discrete rather than continuous.

Aristotle, Leibniz, Newton, and everyone else before Einstein, believed there was a
frame-independent duration between two events. For example, if the time interval
between two lightning flashes is 100 seconds on someone’s accurate clock, then the
interval also is 100 seconds on your own accurate clock, even if you are flying by at an
incredible speed. Einstein rejected this piece of common sense in his 1905 special
theory of relativity when he declared that the time interval between two events
depends on the observer’s reference frame. As Einstein expressed it, “Every reference-
body has its own particular time; unless we are told the reference-body to which the
statement of time refers, there is no meaning in a statement of the time of an event.”
Each reference frame, or reference-body, divides spacetime differently into its time
part and its space part.

In 1908, the mathematician Hermann Minkowski had an original idea in metaphysics


regarding space and time. He was the first person to realize that spacetime is more
fundamental than either time or space alone. As he put it, “Henceforth space by itself,
and time by itself, are doomed to fade away into mere shadows, and only a kind of
union of the two will preserve an independent reality.” The metaphysical assumption
behind Minkowski’s remark is that what is “independently real” is what does not vary
from one reference frame to another. What does not vary is their union, what we now
call “spacetime.” It seems to follow that the division of events into the past ones, the
present ones, and the future ones is also not “independently real.” However, space and
time are not completely equivalent even in relativity theory because time is a
“distinguished” sub-space of the 4-d spacetime continuum. Being distinguished implies
that time is a special dimension unlike the space dimensions, even when we confine
our attention to a single reference frame. For example, a person can move easily
forward and backward in any spatial dimension, but not in the time dimension.

A coordinate system or reference frame is a way of representing space and time using
numbers to represent spacetime points. Science confidently assigns numbers to times
because, in any reference frame, the happens-before order-relation on events is
faithfully reflected in the less-than order-relation on the time numbers (dates) that we
assign to events. In the fundamental theories such as relativity and quantum
mechanics, the values of the time variable t in any reference frame are real numbers,
not merely rational numbers. Each number designates an instant of time, and time is a
linear continuum of these instants ordered by the happens-before relation, similar to
the mathematician’s line segment that is ordered by the less-than relation. Therefore, if
these fundamental theories are correct, then physical time is one-dimensional rather
than two-dimensional, and continuous rather than discrete. These features do not
require time to be linear, however, because a segment of a circle is also a linear
continuum, but there is no evidence for circular time, that is, for causal loops. Causal
loops are worldlines that are closed curves in spacetime.

What about instants? A duration is an ordered set of instants, not a sum of instants.
That is, instants are members of durations, not parts of them. Any duration is infinitely
divisible, and it endlessly divides into more intervals; it never divides into instants. The
parts of durations are just more durations. Instants are like real numbers in that they
are boundaries of durations. They are locations in time, but they are “in” time as
members are in sets, not as parts are in wholes.

The ordering of instants by the happens-before relation, that is, by temporal


precedence, is complete; there are no gaps in the sequence of instants. Knowing an
object, such as an interval of time, is infinitely divisible does not tell you how many
elements or ultimate parts it has, other than that there are infinitely many. It might
have aleph zero or aleph one elements. No physical object is infinitely divisible; and the
reason is that it can be divided into only a finite number of quarks and electrons and
other particles. However, it is often convenient for certain mathematical operations to
treat physical objects as if they were infinitely divisible. Physical space and physical
time are generally believed to be infinitely divisible. Regarding the number of instants
in any (non-zero) duration, time’s being a linear continuum implies the ordered instants
are so densely packed that between any two there is a third, so that no instant has a
next instant. In fact, time’s being a linear continuum implies that there is a
nondenumerable infinity of instants between any two instants, that is, an aleph one
number of instants. There is little doubt that the actual temporal structure of events
can be embedded in the real numbers, but how about the converse? That is, to what
extent is it known that the real numbers can be adequately embedded into the
structure of the instants? The problem here is that, although time is not quantized in
quantum theory, for times shorter than about 10-43 seconds (the so-called Planck time),
science has no experimental grounds for the claim that between any two events there
is a third. Instead, the justification of saying the reals can be embedded into an interval
of instants is that the assumption of continuity is convenient and useful, and that there
are no better theories available.

Because of quantum mechanical considerations, physicists agree that the general


theory of relativity must fail for durations shorter than the Planck time, but they do not
know just how it fails. Most importantly here, there is no agreement among physicists
as to whether the continuum feature of time will be adopted in the future theory of
quantum gravity that will be created to take account of both gravitational and quantum
phenomena. The string theory of quantum gravity predicts that time is continuous, but
an alternative to string theory, loop quantum gravity, does not. (See “Atoms of time.”)

Relativity theory challenges a great many of our intuitive beliefs about time. The theory
is inconsistent with the common sense belief that the temporal order in which two
events occur is independent of the observer’s point of reference. For events occurring
at the same place, relativity theory implies the order is absolute (independent of the
frame) and so agrees with common sense, but for distant events occurring close
enough in time to be in each other’s absolute elsewhere, event A can occur before
event B in one reference frame, but after B in another frame, and simultaneously with
B in yet another frame.

Science impacts our understanding of time in other fundamental ways. Relativity


theory implies there is time dilation between one frame and another. For example, the
faster a clock moves, the slower it runs, relative to stationary clocks. Time dilation
shows itself when a speeding twin returns to find that his (or her) Earth-bound twin has
aged more rapidly. This surprising dilation result has caused some philosophers to
question the consistency of relativity theory by arguing that, if motion is relative, then
we could call the speeding twin “stationary” and it would follow that this twin is now
the one who ages more rapidly. This argument is called the twins paradox of special
relativity. Experts now are agreed that the mistake is within the argument for the
paradox, not within relativity theory. The twins feel different accelerations, so their
motion is not completely symmetric. As is shown in more detail in the Supplement of
Frequently Asked Questions, the argument fails to notice the radically different
relationships that each twin has to the rest of the universe as a whole. This is why one
twin’s proper time is different than the other’s.

[An object's proper time along its worldline, that is, along its path in 4-d spacetime, is
the time elapsed by a clock having the same worldline. Coordinate time is the time
measured by a clock at rest in the (inertial) frame. A clock isn't really measuring the
time in a reference frame other than one fixed to the clock. In other words, a clock
primarily measures the elapsed proper time between events that occur along its own
worldline. Technically, a clock is a device that measures the spacetime interval along
its own worldline. If the clock is at rest in an inertial frame, then it measures the
"coordinate time." If the spacetime has no inertial frame then it can't have a coordinate
time.]

There are two kinds of time dilation. Special relativity’s time dilation involves speed;
general relativity’s also involves acceleration and gravitational fields. Two ideally
synchronized clocks need not stay in synchrony if they undergo different accelerations
or different gravitational forces. This gravitational time dilation would be especially
apparent if one of the two clocks were to fall into a black hole. A black hole can form
when a star exhausts its nuclear fuel and contracts so compactly that the gravitational
force prevents anything from escaping the hole, even light itself. The envelope of no
return surrounding the black hole is its event horizon. As a clock falls toward a black
hole, time slows on approach to the event horizon, and it completely stops at the
horizon (not just at the center of the hole)–relative to time on a clock that remains
safely back on Earth. Every black hole brings an end to time inside itself. If, as many
physicists suspect, the microstructure of spacetime (near the Planck length which is
much smaller than the diameter of a proton) is a quantum foam of changing curvature
of spacetime with black holes forming and dissolving, then time loses its meaning at
this small scale. The philosophical implication is that time exists only when we are
speaking of regions large compared to the Planck length. [If loop quantum gravity turns
out to be the theory that unites quantum mechanics and relativity, then black holes do
not have infinite densities at their center, and light would be trapped inside only for a
finite period, after which what has fallen into the hole will be ejected.]

General Relativity theory may have even more profound implications for time. In 1948,
the logician Kurt discovered radical solutions to Einstein’s equations, solutions in which
there are closed timelike curves, so that as one progresses forward in time along one of
these curves one arrives back at one’s starting point. Gödel drew the conclusion that if
matter is distributed so that there is Gödelian spacetime (that is, with a preponderance
of galaxies rotating in one direction rather than another), then the universe has no
linear time.

b. The Big Bang

The Big Bang is a violent explosion of spacetime that began billions of years ago. It is
not an explosion in spacetime. The Big Bang theory in some form or other is accepted
by the vast majority of astronomers, but it is not as firmly accepted as is the theory of
relativity. Here is a quick story of its origin. In 1922, the Russian physicist Alexander
Friedmann predicted from general relativity that the universe should be expanding. In
1927, the Belgian physicist Georges Lemaitre suggested that galaxy movement could
best be accounted for by this expansion. And in 1929, the American astronomer Edwin
Hubble made careful observations of clusters of galaxies and confirmed that they are
undergoing a universal expansion, on average.

Atoms are not expanding; our solar system is not expanding; even the cluster of
galaxies to which the Milky Way belongs is not expanding. But most every galaxy
cluster is moving away from the others. It is as if the clusters are exploding away from
each other, and in the future they will be very much farther away from each other.
Now, consider the past instead of the future. At any earlier moment the universe was
more compact. Projecting to earlier and earlier times, and assuming that gravitation is
the main force at work, the astronomers now conclude that 13.7 billion years ago the
universe was in a state of nearly zero size and infinite density. Because all substances
cool when they expand, physicists believe the universe itself must have been cooling
down over the last 13.7 billion years, and so it begin expanding when it was extremely
hot. Presently the average temperature of space in all very large regions is 2.7 degrees
Celsius above absolute zero. The Big Bang theory is a theory of how our universe
evolved, how it expanded and cooled from this beginning. This beginning process is
called the “Big Bang.”

As far as we knew back in the 20th century, the entire universe was created in the Big
Bang, and time itself came into existence “at that time.” So, asking what happened
before the Big Bang was properly taken to be like asking what on Earth is north of the
North Pole. With the appearance of the new theories of quantum gravity and the
cosmic landscape in the 21st century, the question has been resurrected as legitimate.

In the literature in both physics and philosophy, descriptions of the Big Bang often
assume that a first event is also a first instant of time and that spacetime did not exist
outside the Big Bang. This intimate linking of a first event with a first time is a
philosophical move, not something demanded by the science. It is not even clear that it
is correct to call the Big Bang an event. The Big Bang “event” is a singularity without
space coordinates, but events normally must have space coordinates. One response to
this problem is to alter the definition of “event” to allow the Big Bang to be an event.
Another response, from James Hartle and Stephen Hawking, is to consider the past
cosmic time-interval to be open rather than closed at t = 0. Looking back to the Big
Bang is then like following the positive real numbers back to ever smaller positive
numbers without ever reaching a smallest positive one. If Hartle and Hawking are
correct that time is actually like this, then the universe had no beginning event. But in
order to simplify the discussion ahead, this article will speak of “the” Big Bang event as
if it were a single origin event.

There are serious difficulties in defending the Big Bang theory’s implications about the
universe’s beginning and its future. Classical Big Bang theory is based on the
assumption that the universal expansion of clusters of galaxies can be projected all the
way back. Yet physicists agree that the projection must fail in the Planck era, that is,
for all times less than 10-43 seconds after “the” Big Bang event. Therefore, current
science cannot speak with confidence about the nature of time within the Planck era. If
a theory of quantum gravity does get confirmed, it should provide information about
this Planck era, and it may even allow physicists to answer the question, “What caused
the Big Bang?”

The scientifically radical, but theologically popular, answer, “God caused the Big Bang,
but He, himself, does not exist in time” is a cryptic answer because it is not based on a
well-justified and detailed theory of who God is, how He caused the Big Bang, and how
He can exist but not be in time. It is also difficult to understand St. Augustine’s remark
that “time itself was made by God.” On the other hand, for a person of faith, belief in
their God is usually stronger than belief in any scientific hypothesis, or in any
epistemological desire for a scientific justification of their remark about God, or in the
importance of satisfying any philosopher’s demand for clarification.

Some physicists are advocating revision of the classical Big Bang theory in order to
allow for the “cosmic landscape” or “multiverse,” in which there are multiple Big Bangs
in parallel universes and an infinite amount of time before our Big Bang. See
(Veneziano, 2006). In some of these universes there is no time dimension. However,
this new theory is not generally accepted by theoretical cosmologists.

c. Infinite Time

There are three ways to interpret the question of whether physical time is infinite: (a)
Was there an infinite amount of time in the universe’s past? (b) Is time infinitely
divisible? (c) Will there be an infinite amount of time in the future?

(a) Was there an infinite amount of time in the past? Aristotle argued “yes,” but by
invoking the radical notion that God is “outside of time,” St. Augustine declared, “Time
itself being part of God’s creation, there was simply no before!” (that is, no time before
God created everything else but Himself). So, for theological reasons, Augustine
declared time had a finite past. After advances in astronomy in the late 19th and early
20th centuries, the question of the age of the universe became a scientific question.
With the acceptance of the classical Big Bang theory, the amount of past time was
judged to be less than 14 billion years because this is when the Big Bang began. The
assumption is that time does not exist independently of the spacetime relations
exhibited by physical events. Recently, however, the classical Big Bang theory has
been challenged. There could be an infinite amount of time in the past according to
some proposed, but as yet untested, theories of quantum gravity based on the
assumptions that general relativity theory fails to hold for infinitesimal volumes. These
theories imply that the beginning of the Big Bang was actually an expansion from a
pre-existing physical state. There was never a singularity. In that case our Big Bang
could be just one bang among other bangs throughout an infinite past of the
landscape. For a discussion of these controversial theories requiring an infinite past
time, see (Veneziano, 2006).

(b) Is time infinitely divisible? Yes, because general relativity and quantum mechanics
require time to be a continuum. But the answer is no if these theories are eventually
replaced by a relativistic quantum mechanics that quantizes time. “Although there
have been suggestions that spacetime may have a discrete structure,” Stephen
Hawking said in 1996, “I see no reason to abandon the continuum theories that have
been so successful.”

(c) Will there be an infinite amount of time in the future? Probably. According to the
classical theory of the Big Bang, the answer depends on whether events will keep
occurring. The best estimate from the cosmologists these days is that the expansion of
the universe is accelerating and will continue forever. There always will be the events
of galaxy clusters getting farther apart, and so future time will have an infinite
duration, even though gravity will continue to compact much of the matter into black
holes.

There have been interesting speculations on how conscious life could continue forever,
despite the fact that the available energy for life will decrease as the universe expands,
and despite the fact that any life swept up into a black hole will reach the center of the
hole in a finite time at which point death will be certain. For an introduction to these
speculations, see (Krauss and Starkman, 2002).

d. Atoms of Time

In the classical theories of relativity and quantum mechanics, time is not quantized, but
is a continuum having the character described above. However, if certain, as yet
untested, theories attempting to unify relativity and quantum mechanics are correct–
such as the theory of loop quantum gravity–then time is composed of discrete
durations lasting about 10-43 second. There is a shortest duration for any possible
event, and time is digital rather than analog.

5. What Kinds of Time Travel are Possible?


Most philosophers believe time travel is possible. In time travel, the traveler’s journey,
as judged by the traveler, takes a different amount of time than the journey does as
judged by those who do not take the journey. That is, there is a difference, and not
merely a verbal disagreement, between the traveler’s inner time or proper time and
the external or coordinate time of those who do not take the journey. However, our
current scientific theories do not allow the external time lapse to be zero; so there is no
“poofing” into the past or “poofing” into the future as in many science fiction stories.

According to relativity theory, there are two ways to travel into another person’s future:
by moving at high speed or by taking advantage of an intense gravitational field. If you
have a fast enough spaceship, you can travel to the year 4,500 C.E. on Earth. You can
affect that future, not just view it. But you can not get back to your own earlier time by
reversing your velocity or reversing the gravitational field. Also, your travel is to
someone else’s future, not your own. You are always in your own present in this sort of
relativistic time travel.
But relativity theory also allows a much stranger kind of time travel, travel to your own
past. For example, in 1949 Kurt Gödel discovered a solution to Einstein’s field
equations that allows continuous, closed future-directed timelike curves. To say this
more simply, Gödel discovered that in some possible worlds that obey the theory of
general relativity, you can eventually arrive into your own past. In this unusual non-
Minkowski spacetime, the universe as a whole is the time machine; no one needs to
build a machine to travel this way. Relativity theory even permits you to travel back
and meet yourself as a child. But, although you can meet yourself, you can not change
what has happened in the past. You can’t go back and prevent Adolf Hitler from gaining
political power in Germany in the 1930s. Despite time travel to the past being
apparently consistent with Einstein’s general theory of relativity, there are several well
known arguments against the physical possibility of travel to the past. Despite the
controversy, none are generally considered to be decisive.

1. If you encountered someone who claimed to be a time traveler, what could you
do to verify the claim? There’s nothing you could do, therefore there will never be
a good reason to believe in time travel.
2. Time travel is impossible because if it were possible we should have seen many
time travelers by now, but nobody has encountered any time travelers.
3. And time travel is impossible because, if there were time travel, then when time
travelers go back and attempt to change history they must always botch their
attempts to change anything, and it will appear to anyone watching them at the
time as if nature is conspiring against them. Since observers have never
witnessed this apparent conspiracy of nature, there is no time travel.
4. If there were travel to the past along a closed timelike curve, then these events
would occur before themselves and after themselves, but this violates our
definition of word “before,” or violates our concept of time, so this odd solution of
Einstein’s equations is not a physically realistic solution.
5. Travel to the past is impossible because it allows the gaining of information for
free. For example, print out this article that you are reading. Enter a time
machine with it. Give me the article before I ever thought about time travel. I
then publish it as this article in this encyclopedia. This all seems to be consistent
with relativity theory, but who first came up with the information in this article?
You had it before I did, but you obtained it from me.
6. Probing the possibility of a contradiction in backwards time travel, the American
philosopher John Earman has described a rocket ship that carries a very special
time machine. The time machine is capable of firing a probe into its own past.
Suppose the ship is programmed to fire the probe on a certain date unless a
safety switch is on. Suppose the safety switch is programmed to be turned on if
and only if the “return” or “impending arrival” of the probe is (or has been)
detected by a sensing device on the ship. Does the probe get launched? It is
launched if and only if it is not launched. The way out of Earman’s paradox seems
to require us to accept that (a) the universe conspires to keep people from
building the probe or the safety switch or an effective sensing device, or (b) time
travel probes must go so far back in time that they never survive and make it
back to the time when they were launched, or (c) time travel into the past is
impossible.

For more discussion of time travel, see the encyclopedia article “Time Travel.”
6. Is the Relational Theory Preferable to the
Absolute Theory?
Absolute theories are theories that imply time exists independently of the spacetime
relations exhibited by physical events. Relational theories imply it does not. Some
absolute theories describe spacetime as being like a container for events. The
container exists with or without events in it. Relational theories imply there is no
container without contents. John Norton’s metaphors might help. Our universe is like a
painting, and absolute spacetime is like the painter’s canvas. If you take away the paint
(the spacetime events) from the painting, you still have the canvas. Relational
spacetime is like citizenship. Take away the citizens (the spacetime events), and you
have no citizenship left.

Everyone agrees time cannot be measured without there being objects and changes,
but the present issue is whether it exists without objects and changes. The absolute or
substantival theories are theories that spacetime could exist even if there were no
physical objects and events in the universe, but relational theories imply that
spacetime is nothing but objects, their events, and the spatiotemporal relationships
among objects and their events, so that spacetime reduces to sets of possible
spatiotemporal relations.

There are two senses of “absolute” that need to be distinguished. As we are using the
term, it means independent of the events. A second sense of “absolute” means
independent of observer or reference frame. Einstein’s theory implies there is no
absolute time in this second sense. Aristotle accepted absolute time in this second
sense, but he rejected it in our sense of being independent of events and took the
relationalist position that, “neither does time exist without change.” [Physics, 218b]

However, the battle lines were most clearly drawn in the early 18th century when
Leibniz argued for the relationalist position against Newton, who had adopted an
absolute theory of time. Leibniz’s principal argument against Newton is a reductio ad
absurdum. Suppose Newton’s absolute space and time were to exist. But one could
then imagine a universe just like ours except with everything shifted five miles east
and five minutes earlier. However, there would be no reason why this shifted universe
does not exist and ours does. Now we have arrived at a contradiction because, if there
is no reason for our universe over the shifted universe, then we have violated Leibniz’s
Principle of Sufficient Reason: that there is an understandable reason for everything
being the way it is. So, Newton’s absolute space and time do not exist. In short, the
trouble with Newton’s absolutism is that it leads to too many unnecessary possibilities.

Newton offered this two-part response: (1) Leibniz is correct to accept the Principle of
Sufficient Reason regarding the rational intelligibility of the universe. But there do not
have to be knowable reasons for humans; God might have had His own sufficient
reason for creating the universe at a given place and time even though mere mortals
cannot comprehend His reasons. (2) The bucket thought-experiment shows that
acceleration relative to absolute space is detectable; thus absolute space is real, and if
absolute space is real, so is absolute time. Suppose we tie a bucket’s handle to a rope
hanging down from a tree branch. Partially fill the bucket with water, and let it come to
equilibrium. Notice that there is no relative motion between the bucket and the water,
and in this case the water surface is flat. Now spin the bucket, and keep doing this until
the angular velocity of the water and the bucket are the same. In this second case
there is also no relative motion between the bucket and the water, but now the water
surface is concave. So spinning makes a difference, but how can a relational theory
explain the difference in the shape of the surface? It can not, says Newton. When the
bucket and water are spinning, what are they spinning relative to? Because we can
disregard the rest of the environment including the tree and rope, says Newton, the
only explanation of the difference in surface shape between the non-spinning case and
the spinning case is that when it is not spinning there is no motion relative to absolute
space, but when it is spinning there is motion relative to space itself, and thus space
itself is acting upon the water surface to make it concave. Alternatively expressed, the
key idea is that the presence of centrifugal force is a sign of rotation relative to
absolute space. Leibniz had no rebuttal. So, for many years thereafter, Newton’s
absolute theory of space and time was generally accepted by European scientists and
philosophers.

One hundred years later, Kant entered the arena on the side of Newton. In a space
containing only a single glove, said Kant, Leibniz could not account for its being a right
glove versus a left glove because all the internal relationships would be the same in
either case. However, we all know that there is a real difference between a right and a
left glove, so this difference can only be due to the glove’s relationship to space itself.
But if there is a “space itself,” then the absolute theory is better than the relational
theory.

Newton’s absolute theory of time was dominant in the 18th and 19th centuries, even
though during those centuries Huygens, Berkeley, and Mach had entered the arena on
the side of Leibniz. In the 20th century, Reichenbach and the early Einstein declared
the special theory of relativity to be a victory for the relational theory. Special relativity,
they said, ruled out a space-filling aether, the leading candidate for absolute space, so
the absolute theory was incorrect. And the response to Newton’s bucket argument is to
note Newton’s error in not considering the environment. Einstein agreed with Mach’s
view of the 19th century that, if you hold the bucket still but spin the background stars,
the water will creep up the side of the bucket. Although it was initially thought by
Einstein and others that relativity theory supported Mach, Lawrence Sklar (Sklar, 1976,
pp. 219-21) argues that this may not be correct.

Many philosophers argue that Reichenbach and the early Einstein have been
overstating the amount of metaphysics that can be extracted from the physics.
Remember the ambiguity in “absolute” mentioned above? There is absolute in the
sense of independent of reference frame and absolute in the sense of independent of
events. Which sense is ruled out when we reject a space-filling aether? The critics
admit that general relativity does show that the curvature of spacetime is affected by
the distribution of matter, so today it is no longer plausible for an absolutist to assert
that the “container” is independent of the matter it contains. But, so they argue,
general relativity does not rule out a more sophisticated absolute theory–to be
discussed below. By the end of the 20th century, absolute theories had gained some
ground thanks to the arguments of John Earman, Michael Friedman, Adolf Grünbaum,
and Tim Maudlin.

In 1969, Sydney Shoemaker presented an argument to convince us of the


understandability of time existing without change, as Newton’s absolutism requires.
Divide space into three disjoint regions, called region 3, region 4, and region 5. In
region 3, change ceases every third year for one year. People in regions 4 and 5 can
verify this and convince the people in region 3 after they come back to life at the end
of their frozen year. Similarly, change ceases in region 4 every fourth year for a year;
and change ceases in region 5 every fifth year. Every sixty years, that is, every 3 x 4 x
5 years, all three regions freeze simultaneously for a year. In year sixty-one, everyone
comes back to life, time having marched on for a year with no change. But
philosophers of time point out that, even if Shoemaker’s scenario shows time’s existing
without change is understandable, the deeper question is whether time does exist
without change.

Here is one argument that it does. Must the relationist say there can be no “empty”
time? If events occur in a room before and after 11:01 AM, but not exactly at 11:01 AM,
must the relationalist say there never was a time of 11:01 AM in the room? To avoid
saying “yes,” which would be absurd, a relationalist might say 11:01 exists in the room
and everywhere else because somewhere outside the room something is happening
then, and somehow or other sense can be made of time in the room in terms of these
external events. The absolutist then asks us to consider the possibility that the room is
the whole universe. In that case, the relationalist response to losing 11:01 AM would
probably be to say possible events occur then in the room even if actual events do not.
But now look where we are, says the absolutist. If the relational theory is going to
consider spacetime points to be permanent possibilities of the location of events, then
the relationalist theory collapses into substantivalism. This is because, to a
substantivalist, a spacetime point is also just a place where something could happen.

Hartry Field offers another argument for the absolute theory by pointing out that
modern physics requires gravitational and electromagnetic fields that cover spacetime–
a light wave, say, is considered to be a ripple in the field. The fields are states of
spacetime, with the field having a value (a number or vector) at points throughout the
field. These fields cannot be states of some Newtonian aether, but there must be
something to have the field properties. What else except substantive spacetime points?

7. Does Time Flow?


“It is as if we were floating on a river, carried by the current past the manifold of events
which is spread out timelessly on the bank,” said one philosopher trying to capture
time’s flow with a helpful metaphor. Santayana offered another: “The essence of
nowness runs like fire along the fuse of time.” The philosopher’s goal is to clarify the
idea of time’s flow, the passage of time. Everyone agrees that the passage of time
“appears” to us humans to flow, although few scientists or philosophers believe that all
conscious beings recognize the flow; hawks do not, although they are apt at spotting
the movements of their prey. Even if time does flow, there is the additional question of
whether the flow can change. Can physical time’s flow be slower on Friday afternoon,
compared to Monday morning?

There are two categories of theories of time’s flow. The first, and most popular among
physicists, is that the flow is an illusion, the product of a faulty metaphor. Time exists,
things change, but time does not flow objectively, although there may well be some
objective feature of our brains that causes us to believe we are experiencing a flow of
time; but in that case time flows only in a subjective sense of the term. The theory is
sometimes characterized as a “myth-of-passage” theory. As we shall see, this theory of
time’s flow is normally the one adopted by those who believe McTaggart’s B-series is
more fundamental than his A-series.

The second category of theories of time’s flow contains theories implying that the flow
is objective, a feature of our mind-independent reality that is to be found in, say, today
scientific laws, or, if it has been missed there, then in future scientific laws. These
theories are called “dynamic theories” of time. This sort of theory of time’s flow is
closer to common sense, and has historically been the more popular theory among
philosophers.

Some dynamic theories imply that the flow is a matter of events changing from being
indeterminate in the future to being determinate in the present and past. Time’s flow is
really events becoming determinate. Thus dynamic theorists speak of time’s flow as
“temporal becoming.” Another dynamic theory implies that the flow is a matter of
events changing from being future, to being present, to being past. This is the kind of
flow associated with McTaggart’s A-series of events.

Opponents of dynamic theories complain that when events change in these senses, the
change is not a real change in the event’s essential, intrinsic properties, but only in the
event’s relationship to the observer. For example, saying the death of Queen Anne is
an event that changes from present to past is no more of a real change in the event
than saying her death changed from being approved of to being disapproved of. This
extrinsic change in approval does not count as a real change in her death, and neither
does the so-called change from present to past. Attacking the notion of time’s flow in
this manner, Grünbaum said: “Events simply are or occur…but they do not ‘advance’
into a pre-existing frame called ‘time.’ …[T]ime is a system of relations between
events, and as events are, so are their relations. An event does not move and neither
do any of its relations.” So, Grünbaum denies the objective nature of McTaggart’s A-
series and points out that the flow of time is an illusion or myth.

Instead of arguing that events change their properties, some advocates of the dynamic
theory of time embrace the flow of time by saying that the flow is reflected in the
change over time of truth values of a sentence or proposition. For example, the
sentence “It is now raining” was true during the rain yesterday but has changed to
false on today’s sunny day. It is these sorts of truth value changes that are at the root
of time’s flow. In response, critics suggest that the indexical (or token reflexive)
sentence “It is now raining” has no truth value because the reference of “now” is
unspecified. If it can not have a truth value, it can not change its truth value. However,
the sentence is related to a sentence that does have a truth value. Supposing it is now
midnight here on April 1, 2007 in Sacramento, California, then the indexical sentence
“It is now raining” is related to the complete or context-explicit sentence “It is raining
at midnight on April 1, 2007 in Sacramento.” Only these non-indexical, non-context-
dependent, complete sentences have truth values, and these truth values do not
change with time. So, events do not change their properties because complete
sentences do not change their truth values.

Other advocates of the dynamic theory of time ask us to analyze time’s flow in terms of
facts that come into existence. This coming into existence of facts, the actualization of
new states of affairs, is time’s flow.
Tim Maudlin argues for a version of the dynamic theory that is very different than all of
the above. He argues that the objective flow of time is fundamental and unanalyzable;
it is a fundamental, irreducible fact that time passes, and this passage just is the flow
of time. He is happy to say “time does indeed pass at the rate of one hour per hour”
(Maudlin, 2007, p. 112), although other philosophers have called this rate
“meaningless.” Maudlin also is an advocate of the block universe theory and believes
the passage of time is an ingredient of this single block entity.

Regardless of how the metaphor of time’s flow is analyzed, or even if it is taken as


fundamentally unanalyzable, the passage of time implies a direction of time.

8. What Gives Time its Direction or “Arrow”?


a. What Needs to be Explained

The arrow of time is what distinguishes a group of events ordered by the happens-
before relation from those ordered by its converse, the happens-after relation. Time’s
arrow is evident in the process of mixing cool cream into hot coffee. You soon get
lukewarm coffee, but you never notice the reverse–lukewarm coffee separating into a
cool part and a hot part. Such is the way this irreversible thermodynamic process goes.
Time’s arrow is also evident when you prick a balloon. The air inside the balloon rushes
out; it never stays in the balloon as it was before the pricking. So, the pricking starts an
irreversible process. The arrow of a physical process is the way it normally goes, the
way it normally unfolds through time. If a process goes only one-way, we call it an
“irreversible process.” (Strictly speaking, a reversible process is one that is reversed by
an infinitesimal change of its surrounding conditions, but we can overlook this fine
point because of the general level of the present discussion.) The amalgamation of the
universe’s irreversible processes produces the cosmic arrow of time, the master arrow.
Usually this arrow is what is meant when one speaks simply of “time’s arrow.” By
convention, we say the arrow is directed toward the future.

There are many goals for a fully developed theory of time’s arrow. It should tell us (1)
why this arrow exists; (2) why the arrow is apparent in macro processes but not micro
processes; (3) what it would be like for the arrow to reverse direction; (4) what the
relationships are among the various more specific arrows of time–the various
temporally asymmetric processes such as entropy increases [the thermodynamic
arrow], causes preceding their effects [the causal arrow], light radiating from its source
rather than converging into it [the electromagnetic arrow], and our knowing the past
more easily than the future [the knowledge arrow]; and (5) what are the characteristics
of a physical theory that pick out a preferred direction in time.

Because the physical processes we commonly observe do have an arrow, you might
think that an inspection of the basic physical laws would readily reveal time’s arrow. It
will not. With very minor exceptions, all the basic laws of fundamental processes are
time symmetric. (It is assumed here that the second law of thermodynamics is not
basic but somehow derived.) This means, according to a principal definition of time
symmetry, that if a certain process is allowed by the laws, then that process reversed
in time is also allowed, and either direction is as probable as the other. Maxwell’s
equations of electromagnetism, for example, can be used to predict that television
signals can exist, but the equations do not tell us whether those signals arrive before or
arrive after they are transmitted. In other words, these basic laws of science do not
imply an arrow of time.

Suppose you have a movie of a basic physical process such as two electrons bouncing
off each other. You can not actually create this movie because the phenomenon is too
small, but forget that fine point for a moment. If you had such a movie, you could run it
forwards or backwards, and both showings would illustrate a possible process
according to the basic laws of science, and they would be equally probable processes.
You could not tell from just looking at the movie whether you were looking at the
original or at it being shown backwards in time. So, time’s arrow is not revealed in this
microscopic process.

The “disappearance” of time’s arrow in microscopic process, does not show that time
itself fades away as you look at briefer and smaller processes; this is because there are
still events happening, and so time still exists there. Also, it is important to note that,
although it is interesting to explain how we humans are able to detect the arrow, the
more challenging philosophical question is to explain why time has an arrow.

b. Explanations or Theories of the Arrow

In the 19th century, the new kinetic theory of gases was supposed to provide the
foundation for all gas behavior, yet this foundational theory is time symmetric. That is,
the theory is insensitive to the arrow of time, to the distinction between past and
future–because a moving molecule could just as well move in one direction as in the
reverse direction. How were the physicists to resolve this apparent contradiction of
having a temporally symmetric theory at the foundation of a theory that is supposed to
account for irreversible gas processes such as the escape of gas from a balloon pricked
with a pin? The first clue was discovered in the mid-19th century by the German
physicist Rudolf Clausius. He devised an early version of the 2nd law of
thermodynamics, which, speaking informally, is the claim that a isolated system will
evolve to be more disordered or complex, with some of its useful energy converting to
heat. [A isolated system is a system left to itself; it is a region isolated from outside
influences, a region where energy can not come in or go out.] That is,

(a) 2nd Law: In an isolated system, entropy never decreases.

Entropy is Clausius’s word for the measure of this disorder; it measures the conversion
of useful to “useless” energy by irreversible processes. As R. A. Fisher expressed it,
entropy changes lead to a progressive disorganization of the physical world, at least
from the human standpoint of the utilization of energy. As time goes on, some sub-
systems do become progressively more organized, such as when we build a house on a
bare lot, but this organization is at the expense of a greater degree of disorganization
elsewhere such as the depletion of natural resources and the digestion of food by the
house builders and, ultimately, the degradation of the sun.

It seemed to many physicists, beginning with Ernst Mach, that time’s arrow–in all
processes and not just in gas behavior–is reducible to or grounded in entropy increase.
This implies that in a universe in maximum equilibrium where entropy changes are
absent, there will not be an arrow of time. This entropy theory of time’s arrow implies
that our having traces of the past but not of the future reduces to entropy increases, as
does our inclination to say causes happen before their effects rather than after.
Another deep question is, “Why should there be more disorder in the future?” The
Austrian physicist Ludwig Boltzmann had an answer in 1872. Boltzmann claimed that it
is a matter of probability because, for complex systems, that is, systems with many
particles, disordered states of the system are more probable than ordered states. There
are many more microstates in which, from a macro perspective, the system is
disordered than microstates in which the system is ordered, so it is very probable that
the system will naturally end up in the most generic possible macrostate. Boltzmann
redefined the concept of entropy in terms of the statistics of molecular motion, and he
deduced a revised 2nd law from probability theory:

(b) 2nd Law: In an isolated system, entropy is likely not to decrease.

His treatment of entropy as being basically a statistical concept was broadly accepted,
as was Mach’s and his claim that time’s arrow is to be explained in terms of entropy
increase.

Boltzmann’s achievement soon had to confront two obstacles, one from Henri Poincaré
and one from Josef Loschmidt. First, Poincaré. A dynamic system is a system defined by
the values of the positions and velocities of all the system’s particles–such as the
places and speeds of the molecules in a cup of coffee. Poincaré’s recurrence theorem
in statistical mechanics says every isolated dynamical system will eventually return to
a state as close to the initial state as we might wish. Wait long enough, and the
lukewarm coffee will separate into hot coffee and cool cream. This reversal would be
expected to take 10N seconds, where N is the number of molecules involved. The
number is staggering, but still finite; so, strictly speaking, there are no irreversible
processes and no long term entropy increase. Whenever entropy rises it will eventually
fall. That implies there is an apparent contradiction between Poincaré’s theorem and
Boltzmann’s.

To avoid this Poincaré problem, physicists redefined the second law:

(c) 2nd Law: In an isolated system, entropy is likely not to decrease for any period of
time that is short compared to the Poincaré period for that system.

Josef Loschmidt pointed out another problem with Boltzmann’s approach to the arrow
of time. Loschmidt realized that Boltzmann’s statistical mechanics predicts for any
point in time not only that entropy should be higher in the future but also that it should
be higher in the past. However, we know that it was not higher in the past. Here is a
graph representing this knowledge.
The conclusion to be drawn from this is that entropy increase is only part of the story of
time’s cosmic arrow.

Loschmidt suggested that the low entropy in the past must be explained by what the
initial conditions happened to be like at the beginning of the universe. Boltzmann
agreed. Among cosmologists, this is now the generally accepted answer to the origin of
time’s arrow.

Yet this answer leads naturally to the request for an explanation of the initial
configuration of our universe. Is this temporally asymmetric initial boundary condition
simply a brute fact, as many physicists believe, or are there as yet undiscovered laws
to explain the fact, as many other physicists believe–either to explain it as necessarily
having had to happen or to explain it as having been highly probable? Objecting to
inexplicable initial facts as being unacceptably ad hoc, the Swiss physicist Walther Ritz
and, more recently, Roger Penrose, say we must not yet have found the true laws (or
invented the best laws) underlying nature’s behavior. We need to keep looking for
basic, time asymmetrical laws in order to account the initial low entropy and thus for
time’s arrow.

The low entropy appears to be due to the microscopic Big Bang region having just the
right amount of homogeneity or smoothness so that galaxies would eventually form. If
it were intially smoother, then there would be no congealing of matter into galaxies; if
it were intially less smooth, then most all the matter would have long ago ended up in
large black holes. So, the issue of how to explain the thermodynamic arrow is the issue
of why the Big Bang region had just the right smoothness.

c. Multiple Arrows
Consider the difference between time’s arrow and time’s arrows. The direction of
entropy change is the thermodynamic arrow. Here are some suggestions for additional
arrows:

1. There are records of the past but not of the future.


2. It is easier to know the past than to know the future.
3. Light and radio waves spread out from, but never converge into, a point.
4. The universe expands rather than shrinks.
5. Causes precede their effects.
6. We see black holes but never white holes.
7. Conscious actions affect the future but not the past.
8. B meson decay, neutral kaon decay, and Higgs boson decay are each different in
a time reversed world.
9. Quantum mechanical measurement collapses the wave function.
10. Possibilities decrease as time goes on.

Most physicists suspect all these arrows are linked so that we can not have some
arrows reversing while others do not. For example, the collapse of the wave function is
generally considered to be due to an increase in the entropy of the universe. However,
the linkage of all the arrows may require as yet undiscovered laws.

d. Reversing Time

But could all the arrows have pointed the other way? That is, could the cosmic arrow of
time have gone the other way? Most physicists suspect that the answer is yes, and it
would have gone the other way if the initial conditions of the universe at the Big Bang
had been different.

Should we also expect that at some time in the future all the arrows will reverse?
Unfortunately, it is still an open question in philosophy as to what it means for time’s
arrow to reverse. For a technical introduction to the debate, see Savitt, pp. 12-19.

Supposing the cosmic arrow of time were to reverse, it would be possible for our past
to be re-created and lived in reverse order. This re-occurrence of the past is different
than the re-living of past events via time travel. With cyclical time or with time travel in
a causal loop, the past is re-visited in the original order that the past events occurred;
the past is not visited in reverse order.

Philosophers have gone on to ask other interesting questions about different scenarios
involving the reversal of time’s arrow. Suppose the cosmic arrow of time were someday
to reverse in a distant, populated region far away from Earth. Imagine what life would
be like for the time-reversed people. First off, would it be possible for them to be
conscious? Assuming consciousness is caused by brain processes, could there be
consciousness if their nerve pulses reversed, or would this reversal destroy
consciousness? This is a difficult question, but supposing the answer is that they would
be conscious, and supposing that anyone’s future is what will happen, not what has
happened, then what would their experience be like? It has been suggested that if we
were able to watch them in their region of space, they would appear to us to be pre-
cognitive. Could they use this to win gambling bets on, say, the roll of the dice?
Probably not, say other philosophers who argue that the inner experience of time-
reversed people must be no different than ours.
If Aristotle were correct that the future, unlike the past, is undetermined or open, then
the future of people in the time-reversed region would be open, too. But it is like our
past. What can we conclude from this? Do we conclude that our past might really be
undetermined and open, too? That our past could change?

And there are other questions. Consider communication between the two regions. If we
sent a signal to the time-reversed region, could our message cross the border, or would
it dissolve there, or would it bounce back? If they successfully sent a recorded film
across the border to us, should we play it in the ordinary way or in reverse? If the arrow
of time were to reverse in some region, would not dead people in that region become
undead, but is that metaphysically possible?

9. Is Only the Present Real?


Have past objects, such as dinosaurs, slipped out of existence? More generally, we are
asking whether the past is real. How about the future? Philosophers are divided into
three camps on the question of the reality of the past, present, and future. The
presentist viewpoint maintains that the past and the future are not real, and that only
the present is real, so if a statement about the past is true, this is because some
present facts make it true. Advocates of a growing past argue that, in addition to the
present, the past is also real. Reality “grows” with the coming into being of determinate
reality from an indeterminate or potential reality. “The world grows by accretion of
facts,” says Richard Jeffrey. Aristotle (in De Interpretatione, chapter 9) and C. D. Broad
advocated a growing-past theory. Parmenides, Duns Scotus and A. N. Prior are
presentists.

Opposing both presentism and the growing past theory, Bertrand Russell, J.J.C. Smart,
W.V.O. Quine, Adolf Grünbaum, and Paul Horwich object to assigning special ontological
status to the present. They say there is no objective ontological difference among the
past, the present, and the future just as there is no ontological difference between here
and there. Yes, we thank goodness that the pain is there rather than here, and past
rather than present, but these differences are subjective, being dependent on our point
of view. This ontology of time is called the block universe theory because it regards
reality as a single block of spacetime with its time slices ordered by the temporally-
before relation. It is mental perspectives only that divide the block into a past part, a
present part, and a future part. The future, by the way, is the actual future, not all
possible futures. William James coined the term “block universe,” but the theory is also
called “eternalism” and the “static theory of time.”

Although presentists say dinosaurs are not real, whereas eternalists say that dinosaurs
are as real as anything in the present, another camp of philosophers argue that the
presentist-eternalist debate is merely verbal because each side is using the word “real”
in a different sense; the presentist uses it in a tensed sense, whereas the eternalist
uses it in an untensed sense.

The presentist and the advocate of the growing past will usually unite in opposition to
the block universe (eternalism) on the grounds that it misses the special “open”
character of the future and the equally significant point that the present is so much
more vivid to a conscious being than is any other time-slice of spacetime. The
advocates of the block universe counter that only the block universe can make sense of
relativity’s implication that, if people are in certain relative motions, an event in person
A’s present can be in person B’s future. Presentism and the growing-past theories must
suppose that this event is both real and unreal because it is real for A but not real for B.
Surely that conclusion is unacceptable, they claim. Their two key assumptions here are
that relativity does provide an accurate account of the spatiotemporal relations among
events, and that if there is some frame of reference in which two events are
simultaneous, then if one of the events is real, so is the other.

Opponents of the block universe charge that it does not provide an accurate account of
the way things are because it leaves out “the now” or “the present.” This metaphysical
dispute was fueled by Einstein who said:

Since there exists in the four dimensional structure no longer any slices which
represent “now” objectively…it appears more natural to think of physical reality as a
four dimensional existence instead of, as hitherto, the evolution of a three dimensional
existence.

Many philosophers, however, do not agree with Einstein.

This philosophical dispute has taken a linguistic turn by focusing upon a question about
language: “Are predictions true or false at the time they are uttered?” Those who
believe in the block universe (and thus in the determinate reality of the future) will
answer “Yes” while a “No” will be given by presentists and advocates of the growing
past. The issue is whether contingent sentences uttered now about future events are
true or false now rather than true or false only in the future at the time the predicted
event is supposed to occur.

Suppose someone says, “Tomorrow the admiral will start a sea battle.” And suppose
that tomorrow the admiral orders a sneak attack on the enemy ships. And suppose that
this action starts a sea battle. Advocates of the block universe argue that, if so, then
the above sentence was true all along. Truth is eternal or fixed, they say, and “is true”
is a tenseless predicate, not one that merely says “is true now.” These philosophers
point favorably to the ancient Greek philosopher Chrysippus who was convinced that a
contingent sentence about the future is true or false, and it can not be any value in
between such as “indeterminate.” Many others, following a suggestion from Aristotle,
argue that the sentence is not true until it can be known to be true, namely at the time
at which the sea battle occurs. The sentence was not true before the battle occurred. In
other words, predictions have no (classical) truth values at the time they are uttered.
Predictions fall into the “truth value gap.” This position that contingent sentences have
no classical truth values is called the Aristotelian position because many researchers
throughout history have taken Aristotle to be holding the position in chapter 9 of On
Interpretation–although today it is not so clear that Aristotle himself held it.

The principal motive for adopting the Aristotelian position arises from the belief that if
sentences about future human actions are now true, then humans are fated (or
determined) to perform those actions, and so humans have no free will. To defend free
will, we must deny truth values to predictions.

The Aristotelian argument against predictions being true or false has been discussed as
much as any in the history of philosophy, and it faces a series of challenges. First, if
there really is no free will, or if free will is compatible with fatalism (or determinism),
then the motivation to deny truth values to predictions is undermined.

Second, if it is true that you will perform an action in the future, it does not follow that
now you will not perform it freely, nor that you are not free to do otherwise, but only
that you will not do otherwise. For more on this point about modal logic, see
Foreknowledge and Free Will.

A third challenge arises from moral discussions about the interests of people who are
as yet unborn. Quine argues that if we have an obligation to conserve the environment
for these people, then we are treating them as being as real as the people around us
now. Only the block universe view can make sense of this treatment.

A fourth challenge, from Quine and others, claims the Aristotelian position wreaks
havoc with the logical system we use to reason and argue with predictions. For
example, here is a deductively valid argument:

There will be a sea battle tomorrow.

If there will be a sea battle tomorrow, then we should wake up the admiral.

So, we should wake up the admiral.

Without the premises in this argument having truth values, that is, being true or false,
we cannot properly assess the argument using the standard of deductive validity
because this standard is about the relationships among truth values of the component
statements. Unfortunately, the Aristotelian position says that some of these
components are neither true nor false, so Aristotle’s position is implausible.

In reaction to this fourth challenge, proponents of the Aristotelian argument claim that
if Quine would embrace tensed propositions and expand his classical logic to a tense
logic, he could avoid those difficulties in assessing the validity of arguments that
involve sentences having future tense.

Quine has claimed that the analysts of our talk involving time should in principle be
able to eliminate the temporal indexical words because their removal is needed for
fixed truth and falsity of our sentences [fixed in the sense of being eternal sentences
whose truth values are not relative because the indicator words have been replaced by
times, places and names, and whose verbs are treated as tenseless], and having fixed
truth values is crucial for the logical system used to clarify science. “To formulate
logical laws in such a way as not to depend thus upon the assumption of fixed truth and
falsity would be decidedly awkward and complicated, and wholly unrewarding,” says
Quine.

Philosophers are still very divided on the issues of whether only the present is real,
what sort of deductive logic to use, and whether future contingent sentences have
truth values.

10. Are there Essentially Tensed Facts?


All the world’s cultures have a conception of time, but in only half the world’s
languages is the ordering of events expressed in the form of tense (Pinker, p. 189). The
English language, for example, expresses conceptions of time with tenses but also with
aspect and with adverbial time phrases such as “now,” “tomorrow” and “twenty-three
days ago.” Philosophers have asked what we are basically committed to when we use
tenses to “locate” an event in the past, in the present, or in the future. For example,
what do we make of the past tense verb in saying, “Mohammed’s birth occurred
centuries ago”? There are two major answers. One answer is that tense distinctions
represent objective features of reality that are not captured by the popular block
universe approach. This answer takes tenses very seriously and is called the tensed
theory of time, or the A-theory in McTaggart’s sense of A vs. B. A second answer to the
question of the significance of tenses is that they are subjective features of the
perspective from which the subject views the universe. Actually this disagreement isn’t
really about tenses in the grammatical sense, but about the significance of the
distinctions of past, present, and future which those tenses are used to mark.

On the tenseless theory of time, or the B-theory, whether the birth of Mohammed
occurred there depends on the speaker’s perspective; similarly, whether the birth
occurs then is equally subjective. The proponent of the tenseless view does not deny
the importance or coherence of talk about the past, but will say it really is (or should be
analyzed as being) talk about our own relation to events. My assertion that
Mohammed’s birth has occurred might be analyzed as asserting that the birth event
happens before the event of my writing this sentence.

This controversy is often presented as a dispute about whether tensed facts exist, with
advocates of the tenseless theory objecting to tensed facts such as the fact of
Mohammed’s having been born. The primary function of tensed facts is to make tensed
sentences true. For the purposes of explaining that point, let us uncritically accept the
Correspondence Theory of Truth and apply it to the following past tense sentence:

Custer died in Montana.

If we apply the Correspondence Theory directly to this sentence, we would say that

The sentence “Custer died in Montana” is true because it corresponds to the tensed
fact that Custer died in Montana.

Opponents of tensed facts argue that the Correspondence Theory should be applied
only indirectly. One approach, the classical tenseless approach, argues that the
Correspondence Theory should be applied only to the result of analyzing away tensed
sentences into equivalent sentences that do not use tenses. They might say that the
sentence “Custer died in Montana” has this equivalent “eternal” sentence:

There is a time t such that Custer dies in Montana at time t, and time t is before the
time of the writing of the sentence “Custer died in Montana” by Dowden in the article
“Time” in The Internet Encyclopedia of Philosophy.

In this analysis, the verb dies is logically tenseless (although grammatically it is present
tensed). Applying the Correspondence Theory to this new sentence yields:
The sentence “Custer died in Montana” is true because it corresponds to the tenseless
fact that there is a time t such that Custer dies in Montana at time t, and time t is
before the time of the utterance (or writing) of the sentence “Custer died in Montana”
by Dowden in the article “Time” in The Internet Encyclopedia of Philosophy

This analysis of tenses without appeal to tensed facts is challenged on the grounds that
it can succeed only for utterances or inscriptions, but a sentence can be true even if
never uttered or written by anyone. There are other challenges. Roderick Chisholm and
A. N. Prior claim that the “is” in the sentence “It is now midnight” is essentially present
tensed because there is no translation using only tenseless verbs. Trying to analyze it
as, say, “There is a time t such that t = midnight” is to miss the essential reference to
the present in the original sentence. The latter sentence is always true, but the original
is not, so the tenseless analysis fails. There is no escape by adding “and t is now”
because this last indexical still needs analysis, and we are starting a vicious regress.

Chisholm and Prior say that true sentences using the temporal indexical terms “now,”
“before now,” and “happened yesterday” are part of the facts of the world that science
should account for, and science fails to do this because it does not recognize them as
being real facts. Science, they say, so far restricts itself to eternal facts, such as in the
Minkowski-like spacetime representation of events. These events are sets of spacetime
points. For such events, the reference to time and place is explicit. A Minkowski
spacetime diagram displays only what happens before what, but not which time is
present time, or past, or future. What is missing from the diagram, say Chisholm and
Prior, is some moving point on the time axis representing the observer’s “now” as time
flows up the diagram.

Earlier, Prior [1959] had argued that after a painful event,

one says, e.g., “Thank goodness that’s over,” and [this]…says something which it is
impossible that any use of a tenseless copula with a date should convey. It certainly
doesn’t mean the same as, e.g., “Thank goodness the date of the conclusion of that
thing is Friday, June 15, 1954,” even if it be said then. (Nor, for that matter, does it
mean “Thank goodness the conclusion of that thing is contemporaneous with this
utterance.” Why should anyone thank goodness for that?).

D. H. Mellor, who advocates a newer subjective theory of tenses, says there’s no


mystery about the meaing of tensed sentences that requires tensed facts or tensed
properties. More specifically, he argues that the truth conditions of any tensed
sentence can be explained without tensed facts even if Chisholm and Prior are correct
that some tensed sentences can not be translated into tenseless ones. Mellor would
say it is not the pastness of the painful event that explains why I say, “Thank goodness
that’s over.” My gladness is explained by my belief that the event is past, plus its being
true that the event is past. In addition, tenseless sentences can be used to explain the
logical relations between tensed sentences: that one tensed sentence implies another,
is inconsistent with yet another, and so forth. And understanding truth conditions and
truth implications is the main thing you know when you understand a declarative
sentence. In other words, the meaning of tensed sentences can be explained without
utilizing tensed properties or tensed facts. Then Ockham’s Razor is applied. If we can
do without essentially tensed facts, then we should say essentially tensed facts do not
exist. To summarize, tensed facts were presumed to be needed to account for the truth
of tensed talk; but the analysis shows that ordinary tenseless facts are adequate. So,
there are no essentially tensed facts, according to Mellor.

11. What is Temporal Logic?


Temporal logic is the representation of information about time by using the methods of
symbolic logic. The classical approach to temporal logic is via tense logic, a formalism
that adds tense operators to an existing system of symbolic logic. The pioneer in the
late 1950s was A. N. Prior. He created a new symbolic logic to describe our use of time
words such as “now,” “happens before,” “afterwards,” “always,” and “sometimes”. The
relationships that propositions have to the past, present, and future help to determine
their truth-value. A proposition, such as “Socrates is sitting down” is allowed to be true
at one time and false at another time.

Prior was the first to appreciate that time concepts are similar in structure to modal
concepts such as “it is possible that” and “it is necessary that,” and so he adapted
modal propositional logic for his tense logic. Dummett and Lemmon also made major,
early contributions to tense logic.

One standard system of tense logic is a variant of the S4.3 system of modal logic. In
this formal tense logic, the usual modal operator “it is possible that” is re-interpreted to
mean “at some past time it was the case that.” Let the letter “P” represent this
operator, and add to the axioms of classical propositional logic the modal-like axiom
P(p v q) iff Pp v Pq. The axiom says that for any two present-tensed propositions p and
q, at some past time it was the case that p or q if and only if either at some past time it
was the case that p or at some past time it was the case that q. The S4.3 system’s key
axiom is the equivalence

Pp & Pq iff P(p & q) v P(p & Pq) v P(q & Pp).

This axiom captures part of our ordinary conception of time as a linear succession of
states of the world. Another axiom might state that if proposition Q is true, then it will
always be true that Q has been true at some time. Prior and others have suggested a
wide variety of axioms for tense logic, but logicians still disagree about what axioms
are needed to make correct beliefs about time be theorems that are logical
consequences of those axioms. Some extension of classical tense logic is definitely
needed in order to express “Q has been true for the past three days.”

The concept of being in the past is usually treated by metaphysicians as a predicate


that assigns properties to events, but in this tense logic the concept is treated as an
operator P upon propositions, and this difference in treatment is objectionable to some
metaphysicians.

The other major approach to temporal logic does not use a tense logic. Instead, it
formalizes temporal reasoning within a first-order logic without modal-like tense
operators. This so-called method of “temporal arguments” adds an additional variable,
a time argument, to any predicate involving time in order to indicate how its
satisfaction depends on time. A predicate such as “is less than seven” does not involve
time, but the predicate “is resting” does. If “x is resting” is represented classically as
R(x), where R is a one-argument predicate, then it would be represented in temporal
logic as R(x,t) and would be interpreted as saying x has property R at time t. R has
been changed to a two-argument predicate by adding a “temporal argument.” The
time variable “t” is treated as a new sort of variable with its own axioms. These axioms
might allow time to be a dense linear ordering without endpoints, or to be even more
like the real numbers.

Occasionally the method of temporal arguments uses a special constant symbol, say
“n”, to denote now, the present time. This helps with the translation of common
temporal statements. For example, the statement that Q has always been true may be
translated into first-order temporal logic as

(For all t)[(t < n) → Q(t)].

Some temporal logics allow sentences to lack a classical truth value. The first person to
give a clear presentation of the implications of treating declarative sentences as being
neither true nor false was the Polish logician Jan Lukasiewicz in 1920. To carry out
Aristotle’s suggestion that future contingent sentences do not yet have truth values, he
developed a three-valued symbolic logic, with all grammatical declarative sentences
having the truth-values of True, False, or else Indeterminate [T, F, or I]. Contingent
sentences about the future, such as Aristotle’s prediction that there will be a sea battle
tomorrow, are assigned an I. Truth tables for the connectives of propositional logic are
redefined to maintain logical consistency and to maximally preserve our intuitions
about truth and falsehood. See (Haack, 1974) for more details about this application of
three-valued logic.

Proper Time, Coordinate Systems,


Lorentz Transformations
This Supplement explains some of the key concepts of the Special Theory of Relativity
(STR). It shows how the predictions of STR differ from classical mechanics in the most
fundamental way. It requires some basic mathematical knowledge.

Table of Contents
1. Proper Time
2. The STR Relationship between Space, Time, and Proper Time
3. Coordinate Systems
4. Cartesian Coordinates for Space
5. Choice of Inertial Reference Frame
6. Operational Specification of Coordinate Systems for Classical Space and Time
7. Operational Specification of Coordinate Systems for STR Space and Time
8. Operationalism
9. Coordinate Transformations and Object Transformations
10. Valid Transformations
11. Velocity Boosts in STR and Classical Mechanics
12. Galilean Transformation of Coordinate System
13. Lorentz Transformation of Coordinate System
14. Time and Space Dilation
15. The Full Special Theory of Relativity
16. References and Further Reading

1. Proper Time
The essence of the Special Theory of Relativity (STR) is that it connects three distinct
quantities to each other: space, time, and proper time. ‘Time’ is also called ‘coordinate
time’ or ‘real time’, to distinguish it from ‘proper time’. Proper time is also called clock
time, or process time. It is a measure of the amount of physical process that a system
undergoes. E.g. proper time for an ordinary mechanical clock is recorded by the
number of rotations of the hands of the clock. Alternatively, we might take a
gyroscope, or a freely spinning wheel, and measure the number of rotations in a given
period. We could also take a chemical process with a natural rate, such as the burning
of a candle, and measure the proportion of candle that is burnt over a given period.

Note that these processes are measured by ‘absolute quantities’: the number of times
a wheel spins on its axis, or the proportion of candle that has burnt. These give
absolute physical quantities, and do not depend upon assigning any coordinate system,
as a numerical representation of space or real time does. The numerical coordinate
systems we use firstly require a choice of measuring units (meters and seconds, for
example). Even more importantly, the measurement of space and real time in STR is
relative to the choice of an inertial frame. This choice is partly arbitrary.

Our numerical representation of proper time also requires a choice of units, and we
adopt the same units as we use for real time (seconds). But the choice of a coordinate
system, based on an inertial frame, does not affect the measurement of proper time.
We will consider the concept of coordinate systems and measuring units shortly.

Proper time can be defined in classical mechanics through cyclic processes that have
natural periods – for instance, pendulum clocks are based on counting the number of
swings of a pendulum. More generally, any natural process in a classical system runs
through a sequence of physical states at a certain absolute rate, and this is the ‘proper
time rate’ for the system.

In classical physics, two identical types of systems (with identical types of internal
construction, and identical initial states) are predicted to have the same proper time
rates. That is, they will run through their physical states in perfect correlation with each
other.

This holds even if two identical systems are in relative constant motion with respect to
each other. For instance, two identical classical clocks would run at the same rate, even
if one is kept stationary in a laboratory, while the other is placed in a spaceship
traveling at high speed.

This invariance principle is fundamental to classical physics, and it means that in


classical physics we can define: Coordinate time = Proper time for all natural systems.
For this reason, the distinction between these two concepts of time was hardly
recognized in classical physics (although Newton did distinguish them conceptually,
regarding ‘real time’ as an absolute temporal flow, and ‘proper time’ as merely a
‘sensible measure’ of real time; see his Scholium).

However, the distinction only gained real significance in the Special Theory of
Relativity, which contradicts classical physics by predicting that the rate of proper time
for a system varies with its velocity, or motion through space. The relationship is very
simple: the faster a system travels through space, the slower its internal processes go.
At the maximum possible speed, the speed of light, c, the internal processes in a
physical system would stop completely. Indeed, for light itself, the rate of proper time
is zero: there is no ‘internal process’ occurring in light. It is as if light is ‘frozen’ in a
specific internal state.

At this point, we should mention that the concept of proper time appears more strongly
in quantum mechanics than in classical mechanics, through the intrinsically ‘wave-like’
nature of quantum particles. In classical physics, single point-particles are simple
things, and do not have any ‘internal state’ that represents proper time, but in
quantum mechanics, the most fundamental particles have an intrinsic proper time,
represented by an internal frequency. This is directly related to the wave-like nature of
quantum particles. For radioactive systems, the rate of radioactive decay is a measure
of proper time. Note that the amount of decay of a substance can be measured in an
absolute sense. For light, treated as a quantum mechanical particle (the photon), the
rate of proper time is zero, and this is because it has no mass. But for quantum
mechanical particles with mass, there is always a finite ‘intrinsic’ proper time rate,
represented by the ‘phase’ of the quantum wave. Classical particles do not have any
correlate of this feature, which is responsible for quantum interference effects and
other non-classical ‘wave-like’ behavior.

2. The STR Relationship between Space, Time, and


Proper Time
STR predicts that motion of a system through space is directly compensated by a
decrease in real internal processes, or proper time rates. Thus, a clock will run fastest
when it is stationary. If we move it about in space, its rate of internal processes will
decrease, and it will run slower than an identical type of stationary clock. The
relationship is precisely specified by the most profound equation of STR, usually called
the metric equation (or line metric equation).

(1)

This applies to the trajectory of any physical system. The quantities involved are:

Dt is the amount of proper time elapsed between two points on the trajectory.Dt is the
amount of real time elapsed between two points on the trajectory.

Dr is the amount of motion through space between two points on the trajectory.

c is the speed of light, and depends on the units we choose for space and time.
The meaning of this equation is illustrated by considering simple trajectories depicted
in a space-time diagram.

Figure 1. Two simple space-time trajectories.

If we start at a initial point on the trajectory of a physical system, and follow it to a later
point, we find that the system has covered a certain amount of physical space, Dr, over
a certain amount of real time, Dt, and has undergone a certain amount of internal
process or proper-time, Dt. As long as we use the same units (seconds) to represent
proper time and real time, these quantities are connected by (1). Proper time intervals
are shown in Figure 1 by blue dots along the trajectories. If these were trajectories of
clocks, for example, then the blue dots would represent seconds ticked off by the clock
mechanism.

In Figure 1, we have chosen to set the speed of light as 1. This is equivalent to using
our normal units for time, i.e. seconds, but choosing the units for space as c meters
(instead of 1 meter), where c is the speed of light in meters per second. This system of
units is often used by physicists for convenience, and it appears to make the quantity c
drop out of the equations, since c = 1. However, it is important to note that c is a
dimensional constant, and even if its numerical value is set equal to 1 by choosing
appropriate units, it is still logically necessary in Equation 1 for the equation to balance
dimensionally. For multiplying an interval of time, Dt, by the quantity c converts from a
temporal quantity into a spatial quantity. Equations of physics, just like ordinary
propositions, can only identify objects or quantities of the same physical kinds with
each other, and the role of c as a dimensional constant remains crucial in Equation 1,
for the identity it states to make any sense.

Trajectories in Figure 1

• Trajectory 1 (green) is for a stationary particle, hence Dr = 0 (it has no motion


through space), and putting this value in Equation 1, we find that: Dt = Dt. For a
stationary particle, the amount of proper time is equal to the amount of
coordinate time.
• Trajectory 2 (red) is for a moving particle, and Dr > 0. We have chosen the
velocity in this example to be: v = c/2, half the speed of light. But: v = Dr/Dt
(distance traveled in the interval of time). Hence: Dr = ½cDt. Putting this value
into Equation 1, we get: c²Dt² = c²Dt²-(½cDt)², or: Dt = Ö(¾)Dt » 0.87Dt. Hence
the amount of proper time is only about 87% of coordinate time. (Even though
this trajectory is very fast, proper time is still only slowed down a little.)
• Trajectory 3 (black) is for a particle moving at the speed of light, with v = c,
giving: Dr = cDt. Putting this in Equation 1, we get: c²Dt² = c²Dt²-(cDt)² = 0.
Hence for a light-like particle, the amount of proper time is equal to 0.

Now from the classical point of view, Equation (1) is a surprise – indeed, it seems
bizarre! For how can mere motion through space directly and precisely affect the rate
of physical processes occurring in a system? We are used to the opposite idea, that
motion through space, by itself, has no intrinsic effect on processes. This is at the heart
of the classical Galilean invariance or symmetry. But STR breaks this rule.

We can compare this situation with classical physics, where (for linear trajectories) we
have two independent equations:

(2.a) Dt = Dt

(2.b) Dr = vDt for some (real numbers)

• Equation (2.a) just means that the rate of proper time in a system is invariant –
and we measure it in the same units as coordinate time, t.
• Equation (2.b) just means that every particle or system has some finite velocity
or speed, v, through space, with v defined by: v = Dr/Dt.

There is no connection here between proper time and spatial motion of the system.

The fact that (2) is replaced by (1) in STR is very peculiar indeed. It means that the rate
of internal process in a system like a clock (whether it is a mechanical, chemical, or
radioactive clock) is automatically connected to the motion of the clock in space. If we
speed up a clock in motion through space, the rate of internal process slows down in a
precise way to compensate for the motion through space.

The great mystery is that there is no apparent mechanism for this effect, called time
dilation. In classical physics, to slow down a clock, we have to apply some force like
friction to its internal mechanism: but in STR, the physical process of a system is
slowed down just by moving it around. This applies equally to all physical processes.
For instance, a radioactive isotope decays more slowly at high speed. And even
animals, including human beings, should age more slowly if they move around at high
speed, giving rise to the ‘Twin’s Paradox’.

In fact, time dilation was already recognized by Lorentz and Poincare, who developed
most of the essential mathematical relationships of STR before Einstein. But Einstein
formulated a more comprehensive theory, and, with important contributions by
Minkowski, he provided an explanation for the effects. The Einstein-Minkowski
explanation appeals to the new concept of a space-time manifold, and interprets
Equation 1 as a kind of ‘geometric’ feature of space-time. This view has been widely
embraced in 20th Century physics. By contrast, Lorentz refused to believe in the
‘geometric’ explanation, and he thought that motion through space has some kind of
‘mechanical’ effect on particles, which causes processes to slow down. While Lorentz’s
view is dismissed by most physicists, some writers have persisted with similar ideas,
and the issues involved in the explanation of Equation 1 continue to be of deep
interest, to philosophers at least.

But before moving on to the explanation, we need to discuss the concepts of


coordinate systems for space and time, which we have been assuming so far without
explanation.

3. Coordinate Systems
In physics we generally assume that space is a three dimensional manifold and time is
a one dimensional continuum. A coordinate system is a way of representing space and
time using numbers to represent points. We assign a set of three numbers, (x,y,z), to
characterize points in space, and one number, t, to characterize a point in time.
Combining these, we have general space-time coordinates: (x,y,z,t). The idea is that
every physical event in the universe has a ‘space-time location’, and a coordinate
system provides a numerical description of the system of these possible ‘locations’.

Classical coordinate systems were used by Descartes, Galileo, Newton, Leibniz, and
other classical physicists to describe space. Classical space is assumed to be a three
dimensional Euclidean manifold. Classical physicists added time coordinates, t, as an
additional parameter to characterize events. The principles behind coordinate systems
seemed very intuitive and natural up until the beginning of the C20th, but things
changed dramatically with the STR. One of Einstein’s first great achievements was to
reexamine the concept of a coordinate system, and to propose a new system suited to
STR, which differs from the system for classical physics. In doing this, Einstein
recognized that the notion of a coordinate system is theory dependant. The classical
system depends on adopting certain physical assumptions of classical physics – for
instance, that clocks do not alter their rates when they are moved about in space. In
STR, some of the laws underpinning these classical assumptions change, and this
changes our very assumptions about how we can measure space and time. To
formulate STR successfully, Einstein could not simply propose a new set of physical
laws within the existing classical framework of ideas about space and time: he had to
simultaneously reformulate the representation of space and time. He did this primarily
by reformulating the rules for assigning coordinate systems for space and time. He
gave a new system of rules suited to the new physical principles of STR, and
reexamined the validity of the old rules of classical physics within this new system.

A key feature Einstein focused on is that a coordinate system involves a system of


operational principles, which connect the features of space and time with physical
processes or ‘operations’ that we can use to measure those features. For instance, the
theory of classical space assumes that there is an intrinsic distance (or length) between
points of space. We may take distance itself to be an underlying feature of ‘empty
space’. Geometric lines can be defined as collections of points in space, and line
segments have intrinsic lengths, prior to any physical objects being placed in space.
But of course, we only measure (or perceive) the underlying structure of space by using
physical objects or physical processes to make measurements. Typically, we use
‘straight rigid rulers’ to measure distances between points of space; or we use
‘uniform, standard clocks’ to measure the time intervals between moments of time.
Rulers and clocks are particular physical objects or processes, and for them to perform
their measurement functions adequately, they must have appropriate physical
properties.

But those physical properties are the subject of the theories of physics themselves.
Classical physics, for example, assumes that ordinary rigid rulers maintain the same
length (or distance between the end-points) when they are moved around in space. It
also assumes that there are certain types of systems (providing ‘idealized clocks’) that
produce cyclic physical processes, and maintain the same temporal intervals between
cycles through time, even if we move these systems around in space.

These assumptions are internally consistent with principles of measurement in classical


physics. But they are contradicted in STR, and Einstein had to reformulate the
operational principles for measuring space and time, in a way that is internally
consistent with the new physical principles of STR.

We will briefly describe these new operational principles shortly, but there are some
features of coordinate systems that are important to appreciate first.

Coordinates as a mathematical language for time and space

The assignment of a numerical coordinate system for time or space is thought of as


providing a mathematical language (using numbers as names) for representing
physical things (time and space). In a sense, this language could be ‘arbitrarily chosen’:
there are no laws about what names can be used to represent things. But naturally
there are features that we want a coordinate system to reflect. In particular, we want
the assignment of numbers to directly reflect the concepts of distance between points
of space, and the size of intervals between moments of time.

We perform mathematical operations on numbers, and we can subtract two numbers to


find the ‘numerical distance’ between them. For numbers are really defined as certain
structures, with features such as continuity, and we want to use the structures of
number systems to represent structural features of space and time.

For instance, we assume in our fundamental physical theory that any two interevals of
time have intrinsic magnitudes, which can be compared to each other. The ‘intrinsic
temporal distance’ between two moments, t1 and t2, may be the same as that between
two quite different moments, t3 and t4. We naturally want to assign numbers to times so
that ordinary numerical subtraction corresponds to the ‘intrinsic temporal distance’
between events. We choose a ‘uniform’ coordinate system for time to achieve this.
Figure 2. A Coordinate system for time gives a mathematical language for a physical
thing.
Numbers are used as names for moments of time.

4. Cartesian Coordinates for Space


Time is simple because it is one-dimensional. Three-dimensional space is much more
complex. Because space is three dimensional, we need three separate real numbers to
represent a single point. Physicists normally choose a Cartesian coordinate system to
represent space. We represent points in this system as: r = (x,y,z), where x, y, and z
are separate numerical coordinates, in three orthogonal (perpendicular) directions.

The numerical structure with real-number points: (x,y,z) is denoted in mathematics


as: . Three dimensional space itself (a physical thing) is denoted as: . A Cartesian
coordinate system is a special kind of mapping between points of these two structures.
It makes the intrinsic spatial distance between two points in E 3 be directly reflected by
the ‘numerical distance’ between their numerical coordinates in .

The numerical distances in are determined by a numerical function for length. A line
from the origin: (0,0,0), to the point r = (x,y,z), which is called the vector r, has its
length given by the Pythagorean formula:

|r| = √(x²+y²+z²).

More generally, for any two points, r1 = (x1, y1, z1), and: r2 = (x2, y2, z2), the distance
function is:

|r2 – r1| = √((x2 – x1)²+ (y2 – y1)²+ (z2 – z1)²)

The special feature of this system is that the lengths of lines in the x, y, or z directions
alone are given directly by the values of the coordinates. E.g. if: r = (x,0,0), then the
vector to r is a line purely in the x-direction, and its length is simply: |r| = x. If r1 =
(x1,0,0), and: r2 = (x2,0,0), then the distance between them is just: |r2 – r1| = (x2 – x1 ).
(As well, a Cartesian coordinate system treats the three directions, x, y, and z, in a
symmetric way: the angles between any pair of these directions is the same, 90 0. For
this reason, a Cartesian system can be rotated, and the same form of the general
distance function is maintained in the rotated system.)

In fact, there are spatial manifolds which do not have any possible Cartesian coordinate
system – e.g. the surface of a sphere, regarded as a two dimensional manifold, cannot
be represented by using Cartesian coordinates. Such spaces were first studied as
geometric systems in the 19th century, and are called non-classical or non-Euclidean
geometries. However, classical space is Euclidean, and by definition:

• Euclidean space can be represented by Cartesian coordinate systems.

We can define alternative, non-Cartesian, coordinate systems for Euclidean space; for
instance, cylindrical and spherical coordinate systems are very useful in physics, and
they use mixtures of linear or radial distance, and angles, as the numbers to specify
points of space. The numerical formulas for distance in these coordinate systems
appear quite different from the Cartesian formula. But they are defined to give the
same results for the distances between physical points. This is the most crucial feature
of the concept of distance in classical physics:

• Distance between points in classical space (or between two events that occur at
the same moment of time) is a physical invariant. It does not change with the
choice of coordinate system.

The form of the numerical equation for distance changes with the choice of coordinate
system; but this is done deliberately to preserve the physical concept of distance.

5. Choice of Inertial Reference Frame


A second crucial concept is the idea of a reference frame. A reference frame specifies
all the trajectories that are regarded as stationary, or at rest in space. This defines the
property of remaining at the same place through time. But the key feature of both
classical mechanics and STR is that no unique reference frame is determined. Any
object that is not accelerating can be regarded as stationary ‘in its own inertial frame’.
It defines a valid reference frame for the whole universe. This is the natural reference
frame ‘from the point of view’ of the object, or ‘relative to the object’. But of course,
there are many possible choices: because given any particular reference frame, any
other frame, defined to give everything a constant velocity relative to the first frame is
also a valid choice.

The class of possible (physically valid) reference frames is objectively determined,


because acceleration is absolutely distinguished from constant motion. Any object that
is not accelerating may be regarded as defining a valid reference frame. But the
specific choice of a reference frame from the range of possibilities is regarded as
arbitrary or conventional. This choice must be made before a coordinate system can be
defined to represent distances in space and time. (Even after we have chosen a
reference frame, there are still innumerable choices of coordinate systems. But the
reference frame settles the definition of distances between events, which must be
defined as the same in any coordinate system relative to a given reference frame.)
The idea of the conventionality of the reference frame is partly evident already in the
choice of a Cartesian coordinate system: for it is an arbitrary matter where we choose
the origin, or point: 0 = (0,0,0), for such a system. It is also arbitrary which directions
we choose for the x, y, and z axes – as long as we make them mutually perpendicular.
We are free to rotate a given set of axes, x, y, z, to produce a new set, x’, y’, and z’,
and this gives another Cartesian coordinate system. Thus, translations and rotations of
Cartesian coordinate systems for space still leave us with Cartesian systems.

But there is a further transformation, which is absolutely central to classical physics,


and involves both time and space. This is the Galilean velocity transformation, or
velocity boost. The essential point is that we need to apply a spatial coordinate system
through time. In pure classical geometry, we do not have to take time into account: we
just assign a single coordinate system, at a single moment of time. But in physics we
need to apply a coordinate system for space at different moments of time. How do we
know whether the coordinate system we apply at one moment of time represents the
same coordinate system we use at a later moment of time?

The principles of classical physics mean that we cannot measure ‘absolute location in
space’ across time. The reason is the fundamental classical principle that the laws of
nature do not distinguish between two inertial frames moving relative to each other at
a constant speed. This is the classical Galilean principle of ‘relativity of motion’.
Roughly stated, this means that uniform motion through space has no effect on
physical processes. And if motion in itself does not affect processes, then we cannot
use processes to detect motion.

Newton believed that the classical conception of space requires there to be absolute
spatial locations through time nonetheless, and that some special coordinate systems
or physical objects will indeed be at ‘absolute rest’ in space. But in the context of
classical physics, it is impossible to measure whether any object is at absolute rest, or
is in uniform motion in space. Because of this, Leibniz denied that classical physics
requires any concept of absolute position in space, and argued that only the notion of
‘relative’ or ‘relational’ space’ is required. In this view, only the relative positions of
objects w.r.t each other are considered real. For Newton, the impossibility of measuring
absolute space does not prevent it from being a viable concept, and even a logically
necessary concept. There is still no general agreement about this debate between
‘absolute’ and ‘relative’ or ‘relational’ conceptions of space. It is one of the great
historical debates in the philosophy of both classical and relativistic physics. However,
it is generally accepted that classical physics makes absolute space undetectable. This
means, at least, that in the context of classical physics there is no way of giving an
operational procedure for determining absolute position (or absolute rest) through
time.

However absolute acceleration is detectable. Accelerations are always accompanied by


forces. This means that we can certainly specify the class of coordinate systems which
are in uniform motion, or which do not accelerate. These special systems are called
inertial systems, or inertial frames, or Galilean frames. The existence of inertial frames
is a fundamental assumption of classical physics. It is also fundamental in STR, and the
notion of an inertial frame is very similar in both theories.

The laws of classical physics are therefore specified for inertial coordinate systems.
They are equally valid in any inertial frame. The same holds for the laws of STR.
However, the laws for transforming from one inertial frame to another are different for
the two theories. To see how this works, we now consider the operational specification
of coordinate systems.

6. Operational Specification of Coordinate Systems


for Classical Space and Time
In classical physics, we can define an ‘operational’ measuring system, which allows us
to assign coordinates to events in space and time.

Classical Time. We imagine measuring time by making a number of uniform clocks,


synchronizing them at some initial moment, checking that they all run at exactly the
same rates (proper time rates), and then moving clocks to different points of space,
where we keep them ‘stationary’ in a chosen inertial frame. We subsequently measure
the times of events that occur at the various places, as recorded by the different clocks
at those places.

Of course, we cannot assume that our system of clocks is truly stationary. The entire
system of clocks placed in uniform motion would also define a valid inertial frame. But
the laws of classical physics mean that clocks in uniform inertial motion run at exactly
the same rates, and so the times recoded for specific events turn out to be exactly the
same, on the assumptions of the classical theory, for any such system of clocks.

Classical Space. We imagine measuring space by constructing a set of rigid


measuring rods or rulers of the same length, which we can (imaginatively at least) set
up as a grid across space, in an inertial frame. We keep all the rulers stationary relative
to each other, and we use them to measure the distances between various events.
Again, the main complication is that we cannot determine any absolutely stationary
frame for the grid of rulers, and we can set up an alternative system of rulers which is
in relative motion. This results in assigning different ‘absolute velocities’ to objects, as
measured in two different frames. However, on the assumptions of the classical theory,
the relative distances between any two objects or events, taken at any given moment
of time, is measured to be the same in any inertial frame. This is because, in classical
physics, uniform motion in itself does not alter the lengths of material objects, or the
forces between systems of objects. (Accelerations do alter lengths).

7. Operational Specification of Coordinate Systems


for STR Space and Time
In STR, the situation is in many ways very similar to classical physics: there is still a
special concept of inertial frames, acceleration is absolutely detectable, and uniform
velocity is undetectable. According to STR, the laws of physics still are invariant w.r.t.
uniform motion in space, very much like the classical laws.

We also specify operational definitions of inertial coordinate systems in STR in a similar


way to classical physics. However, the system sketched above for assigning classical
coordinates fails, because it is inconsistent with the physical principles of STR. Einstein
was forced to reconstruct the classical system of measurement, to obtain a system
which is internally consistent with STR.
STR Time. In STR, we can still make uniform clocks, which run at the same rates when
they are held stationary relative to each other. But now there is a problem
synchronizing them at different points of space. We can start them off synchronized at
a particular common point; but moving them to different points of space already upsets
their synchronization, according to Equation 1.

However, while synchronizing distant clocks is a problem, they nonetheless run at the
same intrinsic rates as each other when held in the same inertial frame. And we can
ensure two clocks are in a common inertial frame as long as we can ensure that they
maintain the same distance from each other. We see how to do this next.

Given we have two clocks maintained at the same distance from each other, Einstein
showed that there is indeed a simple operational procedure to establish
synchronization. We send a light signal from Clock 1 to Clock 2, and reflect it back to
Clock 1. We record the time it was sent on Clock 1 as t 0, and the time it was received
again as a later time, t2. We also record the time it was received at Clock 2 as t 1’ on
Clock 2. Now symmetry of the situation requires that, in the inertial frame of Clock 1,
we must assume that the light signal reached Clock 2 at a moment halfway between t0
and t1, i.e. at the time: t1 = ½(t2 – t0). This is because, by symmetry, the light signal
must take equal time traveling in either direction between the clocks, given that they
are kept at a constant distance throughout the process, and they do not accelerate. (If
the light signal took longer to travel one way than the other, then light would have to
move at different speeds in different directions, which contradicts STR).

Hence, we must resynchronize Clock 2 to make: t1’ = t1. We simply set the hands on
Clock 2 forwards by: (t1 – t1’), i.e. by: ½(t2 – t0) – t1’. (Hence, the coordinate time on
Clock 2 at t1’ is changed to: t1’ + (½(t2 – t0) – t1’) = ½(t2 – t0) = t1.)

This is sometimes called the ‘clock synchronization convention’, and some philosophers
have argued about whether it is justified. But there is no real dispute that this
successfully defines the only system for assigning simultaneity in time, in the chosen
reference frame, which is consistent with STR.

Some deeper issues arise over the notion of simultaneity that it seems to involve. From
the point of view of Clock 1, the moment recorded at: t1 = ½(t2 – t0) must be judged as
‘simultaneous’ with the moment recorded at t1’ on Clock 2. But in a different inertial
frame, the natural coordinate system will alter the apparent simultaneity of these two
events, so that simultaneity itself is not ‘objective’ in STR, except relative to a choice of
inertial frame. We will consider this later.

STR Space. In STR, we can measure space in a very similar way as in classical physics.
We imagine constructing a set of rigid measuring rods or rulers, which are checked to
be the same length in the inertial frame of Clock 1, and we extend this out into a grid
across space. We have to move the rulers around to start with, but when we have set
up the grid, we keep them all stationary in the chosen inertial frame of Clock 1.

We then use this grid of stationary measuring rods to measure the distances between
various events. The main assumption is that identical types of measuring rods (which
are the same lengths when we originally compare them at rest with Clock 1), maintain
the same lengths after being moved to different places (and being made stationary
again w.r.t. Clock 1). This feature is required by STR.
The main complication, once again, is that we cannot determine any absolutely
stationary frame for the grid of rulers. We can set up an alternative system of rulers,
which are all in relative motion in a different inertial frame. As in classical physics, this
results in assigning different ‘absolute velocities’ to most trajectories in the two
different frames. But in this case there is a deeper difference: on the assumptions of
STR, the lengths of measuring rods alter according to their velocities. This is called
space dilation, and it is the counterpart of time dilation.

Nonetheless, Einstein showed that perfectly sensible operational definitions of


coordinate measurements for length, as well as time, are available in STR. But both
simultaneity and length become relative to specified inertial frames.

It is this confusing conceptual problem, which involves the theory dependence of


measurement, that Einstein first managed to unravel, as the prelude to showing how to
radically reconstruct classical physics.

8. Operationalism
Unraveling this problem requires us to specify ‘operational principles’ of measurement,
but this does not require us to embrace an operational theory of meaning. The latter is
a form of positivism, and it holds that the meaning of ‘time’ or ‘space’ in physics is
determined entirely by specifying the procedures for measuring time or space. This
theory is generally rejected by philosophers and logicians, and it was rejected by
Einstein himself in his mature work. According to operationalism, STR changes the
meanings of the concepts of space and time from the classical conception. However,
many philosophers would argue that ‘time’ and ‘space’ have a meaning for us which is
essentially the same as for Galileo and Newton, because we identify the same kinds of
things as time and space; but relativity theory has altered our scientific beliefs about
these things – just as the discovery that water is H2O has altered our understanding of
the nature of water, without necessarily altering the meaning of the term ‘water’. This
semantic dispute is ongoing in the philosophy of science. Having clarified these basic
ideas of coordinate systems and inertial frames, we now turn back to the notion of
transformations between coordinate systems for different inertial frames.

9. Coordinate Transformations and Object


Transformations
Physics uses two different concepts of transformations. It is important to distinguish
these carefully.

• Coordinate transformations: First is the notion of taking the description of a given


process (such as a trajectory), described in one coordinate system, and
transforming to its description in an alternative coordinate system.
• Object transformations: Second is the notion of taking a given process, described
in a given coordinate system, and transforming it into a different process,
described in the same coordinate system as the original process.

The difference is illustrated in the following diagram for the simplest kind of
transformation, translation of space.
Figure 3. Object, Coordinate, and Combined Transformations.

• The transformations in Figure 3 are simple space translations.


• Figure 3 (B) shows an object transformation. The original trajectory (A) is moved
in space to the right, by 4 units. The new coordinates are related to the original
coordinates by: xnew particle ® xoriginal particle + 4.
• Figure 3 (C) shows a coordinate transformation: the coordinate system is moved
to the right by 4 units. The new coordinate system, x’, is related to the original
system, x, by: x’original particle = xoriginal particle + 4. The result ‘looks’ the same as (B).
• Figure 3 (D) shows a combination of the object transformation (B) and a
coordinate transformation, which is the inverse of that in (C), defined by: x’’original
particle = xoriginal particle – 4. The result of this looks the same as the original trajectory
in (A), because the coordinate transformation appears to ‘undo’ the effect of the
object transformation.

10. Valid Transformations


There is an intimate connection between these two kinds of transformations. This
connection provides the major conceptual apparatus of modern physics, through the
concept of physical symmetries, or invariance principles, and valid transformations.

The deepest features of laws or theories of physics are reflected in their symmetry
properties, which are also called invariances under symmetry transformations. Laws or
theories can be understood as describing classes of physical processes. Physical
processes that conform to a theory are valid physical processes of that theory. Of
course, not all (logically) possible processes that we can imagine are valid physical
processes of a given theory. Otherwise the theory would encompass all possible
processes, and tell us nothing about what is physically possible, as opposed to what is
logically conceivable.

Symmetries of a theory are described by transformations that preserve valid processes


of the theory. For instance, time translation is a symmetry of almost all theories. This
means that if we take a valid process, and transform it, intact, to an earlier or later
time, we still have a valid process. This is equivalent to simply setting the ‘temporal
origin’ of the process to a later or earlier time.

Other common symmetries are:

• Rotations in space (if we take a valid process, and rotate it to another direction in
space, we end up with another valid process).
• Translations in space (if we take a valid process, and move it to another position
in space, we end up with another valid process).
• Velocity transformations (if we take a valid process, and give it uniform velocity
boost in some direction in space, we end up with another valid process).

These symmetries are valid both in classical physics and in STR. In classical physics,
they are called Galilean symmetries or transformations. In STR they are called Lorentz
transformations. However, although the symmetries are very similar in both theories,
the Lorentz transformations in STR involve features that are not evident in the classical
theory. In fact, this difference only emerges for velocity boosts. Translations and
rotations are identical in both theories. This is essentially because velocity boosts in
STR involve transformations of the connection between proper time and ordinary space
and time, which does not appear in classical theory.

The concept of valid coordinate transformations follows directly from that of valid
object transformations. The point is that when we make an object transformation, we
begin with a description of a process in a coordinate system, and end up with another
description, of a different process, given in the same coordinate system. Now instead of
transforming the processes involved, we can do the inverse, and make a
transformation of the coordinate system, so that we end up with a new coordinate
description of the original process, which looks exactly the same as the description of
the transformed process in the original coordinate system.

This gives an alternative way of regarding the process, and its transformed image:
instead of taking them as two different processes, we can take them as two different
coordinate descriptions of the same process.
This is connected to the idea that certain aspects of the coordinate system are
arbitrary or conventional. For instance, the choice of a particular origin for time or
space is regarded as conventional: we can move the origins in our coordinate
description, and we still have a valid system. This is only possible because the
corresponding object transformations (time and space translations) are valid physical
transformations.

Physicists tend to regard coordinate transformations and valid object transformations


interchangeably and somewhat ambiguously, and the distinction between the two is
often blurred in applied physics. While this doesn’t cause practical problems, it is
important when learning the concepts of the theory to distinguish the two kinds of
transformations clearly.

11. Velocity Boosts in STR and Classical Mechanics


STR and classical mechanics have exactly the same symmetries under translations of
time and space, and rotations of space. They also both have symmetries under velocity
boosts: both theories hold that, if we take a valid physical process, and give it a
uniform additional velocity in some direction, we end with another valid physical
process. But the transformation of space and time coordinates, and of proper time, are
different for the two theories under a velocity boost. In classical physics, it is called a
Galilean transformation, while for STR it is called a Lorentz transformation.

To see how the difference appears, we can take a stationary trajectory, and consider
what happens when we apply a velocity boost in either theory.

Figure 4. Classical and STR Velocity Boosts give different results.

In both diagrams, the green line is the original trajectory of a stationary particle, and it
looks exactly the same in STR and classical mechanics. Proper time events (marked in
blue) are equally spaced with the coordinate time intervals in both cases.
If we transform the classical trajectory by giving the particle a velocity (in this example,
v = c/2) towards the right, the result (red line) is very simple: the proper time events
remain equally spaced with coordinate time intervals. The same sequence of proper
time events takes the same amount of coordinate time to complete. The classical
particle moves a distance: Dx = v.Dt to the right, where Dt is the coordinate time
duration of the original process.

But when we transform the STR particle, a strange thing happens: the proper time
events become more widely spaced than the coordinate time intervals, and the same
sequence of proper time events takes more coordinate time to complete. The STR
particle moves a distance: Dx’ = v.Dt’ to the right, where: Dt’ > Dt, and hence: Dx’ >
Dx.

The transformations of the coordinates of the (proper time) points of the original
processes are shown in the following table.

Table 1. Example of Velocity Transformation.

We can work out the general formula for the STR transformations of t’ and x’ in this
example by using Equation 1. This requires finding a formula for the transformation of
time-space coordinates:

(t, 0) ® (t’, x’)

We obtain this by applying Equation 1 in the (t’,x’) coordinate system, giving:

(1’)

It is crucial that this equation retains the same form under the Lorentz equation. In this
special case, we have the additional facts that:

(i) Dt = Dt, and:(ii) Dx’ = vDt’


We substitute (i) and (ii) in (1’) to get:

This rearranges to give:

and:

We can see that: Dx’/Dt’ = v. This is a special case of a Lorentz transformation for this
simplest kind of trajectory. Note that if we think of this as a coordinate transformation
which generates the appearance of this object transformation, we need to move the
new coordinate system in the opposite direction to the motion of the object. I.e. if we
define a new coordinate system, (x’,t’), moving at –v (i.e. to the left) w.r.t. the original
(x,t) system, then the original trajectory (which appeared stationary in (x,t)) will appear
to be moving with velocity +v (to the left) in (x’,t’). In general, object transformations
correspond the inverse coordinate transformations.

12. Lorentz Transformations for Velocity Boost V in


the x-direction
The previous transformations is only for points on the special line where: x = 0. More
generally, we want to work out the formulae for transforming points anywhere in the
coordinate system:

(t, x) ® (t’, x’)

The classical formulas are Galilean transformations, and they are very simple.

Galilean Velocity Boost:

(t, x) ® (t, x+vt)t’ = t

x’ = x+vt

The STR formulas are more general Lorentz transformations. The Galilean
transformation is simple because time coordinates are unchanged, so that: t = t’. This
means that simultaneity in time in classical physics is absolute: it does not depend
upon the choice of coordinate system. We also have that distance between two points
at a given moment of time is invariant, because if: x 2 -x1 = Dx, then: x’2 -x’1 = (x2+vt) –
(x1-vt) = Dx. Ordinary distance in space is the crucial invariant quantity in classical
physics.

But in STR, we have a complex interdependence of time and space coordinates. This is
seen because the transformation formulas for both t’ and x’ are functions of both x and
t. I.e. there are functions f and g such that:

t’ = f(x,t) and: x’ = g(x,t)


These functions represent the Lorentz transformations. To give stationary objects a
velocity V in the x-direction, these general functions are found to be:

Lorentz Transformations: and:

The factor: is called γ, letting us write these equations more simply as:

Lorentz Transformations: t’ = γ(t+Vx/c2) and: x’ = γ(x+Vt)

We can equally consider the corresponding coordinate transformation, which would


generate the appearance of this object transformation in a new coordinate system. It is
essentially the same as the object transformation – except it must go in the opposite
direction. For the object transformation, which increases the velocity of stationary
particles by the speed V in the x direction, corresponds to moving the coordinate
system in the opposite direction. I.e. if we define a new coordinate system, and call it
(x’,t’), and place this in motion with a speed –V (i.e. V in the negative-x-direction),
relative to the (x,t) coordinate system, then the original stationary trajectories in (x,t)-
coordinates will appear to have speed V in the new (x’,t’) coordinates.

Because the Lorentz transformation of processes leaves us with valid STR processes,
the Lorentz transformation of a STR coordinate system leaves us with a valid
coordinate system. In particular, the form of Equation 1 is preserved by the Lorentz
transformation, so that we get: . This can be checked by substituting
the formulas for t’ and x’ back into this equation, and simplifying; the resulting
equation turns out to be identical to Equation 1.

13. Galilean Transformation of Coordinate System


One useful way to visualize the effect of a transformation is to make an ordinary space-
time diagram, with the space and time axes drawn perpendicular to each other as
usual, and then to draw the new set of coordinates on this diagram. In these diagrams,
the space axes represent points which are measured to have the same time
coordinates, and similarly, the time axes represent points which are measured to have
the same space coordinates. When we make a velocity boost, these lines of
simultaneity and same-position are altered.

This is shown first for a Galilean velocity boost, where in fact the lines of simultaneity
remain the same, but the lines representing position are rotated:
Figure 5. Galilean Velocity Boost.

• In Figure 5, the (green) horizontal lines are lines of absolute simultaneity. They
have the same coordinates in both t and t’.
• The (blue) vertical lines are lines with the same x-coordinates.
• The (gray) slanted lines are lines with the same x’-coordinates.
• The spacing of the x’ coordinates is the same as the x coordinates, which means
that relative distances between points are not affected.
• The solid black arrow represents a stationary trajectory in (x,t).
• An object transformation of +V moves it onto the green arrow, with velocity: v =
c/2 in the (x,t)-system.
• A coordinate transformation of +V, to a system (x’,t’) moving at +V w.r.t. (x,t),
makes this green arrow appears stationary in the (x’,t’) system.
• This coordinate transformation makes the black arrow appear to be moving at –V
in (x’,t’) coordinates.

14. Lorentz Transformation of Coordinate System


In a Lorentz velocity boost, the time and space axes are both rotated, and the spacing
is also changed.
Figure 6. Rotation of Space and Time Coordinate Axes by a Lorentz Velocity Boost.
Some proper time events are marked in blue.

To obtain the (x’,t’)-coordinates of a point defined in (x,t)-coordinates, we start at the


point, and: (i) move parallel to the green lines, to find the intersection with the (red) t’-
axis, which is marked with the x’-coordinates; and: (ii) move parallel to the red lines, to
find the intersection with the (green) x’-axis, which is marked with the t’-coordinates.
The effects of this transformation on a solid rod or ruler extending from x=0 to x=1,
and stationary in (x,t), is shown in more detail below.
Figure 7. Lorentz Velocity Boost. Magnified view of Figure 6 shows time and space
dilation. The gray rectangle represents a unit of the space-time path of a rod (Rod 1)
stationary in (x,t). The dark green lines represent a Lorentz (object) transformation of
this trajectory, which is a second rod (Rod 2) moving at V in (x,t) coordinates. This is a
unit of the space-time path of a stationary rod in (x’,t’).

15. Time and Space Dilation


Figure 7 shows how both time and space dilation effects work. To see this clearly, we
need to consider the volumes of space-time that an object like a rod traces out.

• The (gray) rectangle PQRS represents a space-time volume, for a stationary rod
or ruler in the original frame. It is 1-meter long in original coordinates (Dx = 1),
and is shown over 1 unit of proper time, which corresponds to one unit of
coordinate time (Dt = 1).
• The rectangle PQ’R’S’ (green edges) represents a second space-time volume, for
a rod which appears to be moving in the original frame. This is how the space-
time volume of the first rod transforms under a Lorentz transformation.
• We may interpret the transformation as either: (i) a Lorentz velocity boost of the
rod by velocity +V (object transformation), or equally: (ii) a Lorentz
transformation to a new coordinate system, (x’,t’), moving at –V w.r.t. (x,t). Note
that:
• The length of the moving rod measured in x is now shorter than the stationary
rod: Dx = 1/γ. This is space dilation.
• The coordinate time between proper time events on the moving rod measured in
t is now longer than for the stationary rod (Dt = γ). This is time dilation.

The need to fix the new coordinate system in this way can be worked out by
considering the moving rod from the point of view of its own inertial system.

• As viewed in its own inertial coordinate system, the green rectangle PQ’R’S’
appears as the space-time boundary for a stationary rod. In this frame:
• PS’ appears stationary: it is a line where: x’ = 0.
• PQ’ appears as a line of simultaneity, i.e. it is a line where: t’=0.
• R’S’ is also a line of simultaneity in t’.
• Points on R’S’ must have the time coordinate: t’=1, since it is at the time t’ when
one unit of proper time has elapsed, and for the stationary object, Dt’ = Dt.
• The length of PQ’ must be one unit in x’, since the moving rod appears the same
length in its own inertial frame as the original stationary rod did.

Time and space dilation are often referred to as ‘perspective effects’ in discussions of
STR. Objects and processes are said to ‘look’ shorter or longer when viewed in one
inertial frame rather than in another. It is common to regard this effect as a purely
‘conventional’ feature, which merely reflects a conventional choice of reference frame.
But this is rather misleading, because time and space dilation are very real physical
effects, and they lead to completely different types of physical predictions than
classical physics.

However, the symmetrical properties of the Lorentz transformation makes it impossible


to use these features to tell whether one frame is ‘really moving’ and another is ‘really
stationary’. For instance, if objects get shorter when they are placed in motion, then
why don’t we simply measure how long objects are, and use this to determine whether
they are ‘really stationary’? The details in Figure 7 reveal why this does not work: the
space dilation effect is reversed when we change reference frames. That is:

• Measured in Frame 1, i.e. in (x,t)-coordinates, the stationary object (Rod 1)


appears longer than the moving object (Rod 2). But:
• Measured in Frame 2, using (x’,t’)-coordinates, the moving object (Rod 2)
appears stationary, while the originally stationary object (Rod 1) moves. But now
the space dilation effect appears reversed, and Rod 2 appears longer than Rod 1!

The reason this is not a real paradox or inconsistency can be seen from the point of
view of Frame 2, because now Rod 1 at the moment of time t’ = 0 stretches from the
point P to Q’’, rather than from P to Q, as in Frame 1. The line of simultaneity alters in
the new frame, so that we measure the distance between a different pair of space-time
events. And PQ’’ is now found to be shorter than PQ’, which is the length of Rod 2 in
Frame 2.

There is no answer, within STR, as to which rod ‘really gets shorter’. Similarly there is
no answer as to which rod ‘really has faster proper time’ – when we switch to Frame 2,
we find that Rod 2 has a faster rate of proper time w.r.t. coordinate time, reversing the
time dilation effect apparent in Frame 1. In this sense, we could consider these effects
a matter of ‘perspective’ – although it is more accurate to say that in STR, in its usual
interpretation, there are simply no facts about absolute length, or absolute time, or
absolute simultaneity, at all.

However, this does not mean that time and space dilation are not real effects. They are
displayed in other situations where there is no ambiguity. One example is the twins’
paradox, where proper time slows down in an absolute way for a moving twin. And
there are equally real physical effects resulting from space dilation. It is just that these
effects cannot be used to determine an absolute frame of rest.

16. The Full Special Theory of Relativity


So far, we have only examined the most basic part of STR: the valid STR
transformations for space, time, and proper time, and the way these three quantities
are connected together. This is the most fundamental part of the theory. It represents
relativistic kinematics. It already has very powerful implications. But the fully
developed theory is far more extensive: it results from Einstein’s idea that the Lorentz
transformations represent a universal invariance, applicable to all physics. Einstein
formulated this in 1905: “The laws of physics are invariant under Lorentz
transformations (when going from one inertial system to another arbitrarily chosen
inertial system)”. Adopting this general principle, he explored the ramifications for the
concepts of mass, energy, momentum, and force.

The most famous result is Einstein’s equation for energy: E = mc². This involves the
extension of the Lorentz transformation to mass. Einstein found that when we Lorentz
transform a stationary particle with original rest-mass m0, to set it in motion with a
velocity V, we cannot regard it as maintaining the same total mass. Instead, its mass
becomes larger: m = γm0, with γ defined as above. This is another deep contradiction
with classical physics.

Einstein showed that this requires us to reformulate our concept of energy. In classical
physics, kinetic energy is given by: E = ½ mv². In STR, there is a more general
definition of energy, as: E = mc². A stationary particle then has a basic ‘rest mass
energy’ of m0c². When it is set in motion, its energy is increased purely by the increase
in mass, and this is kinetic energy. So we find in STR that:

Kinetic Energy = mc²-m0c² = (γ-1)m0c²

For low velocities, with: v << c, it is easily shown that: (γ-1)c² is very close to ½v², so
this corresponds to the classical result in the classical limit of low energies. But for high
energies, the behavior of particles is very different. The discovery that there is an
underlying energy of m0c² simply from rest-mass is what made nuclear reactors and
nuclear bombs possible: they convert tiny amounts of rest mass into vast amounts of
thermal energy.

The main application Einstein explored first was the theory of electromagnetism, and
his most famous paper, in which he defined STR in 1905, is called “Electrodynamics of
Moving Bodies”. In fact, Lorentz, Poincare and others already knew that they needed to
apply the Lorentz transformation to Maxwell’s theory of classical electromagnetism,
and had succeeded a few years earlier in formulating a theory which is extremely
similar to Einstein’s in its predictions. Some important experimental verification of this
was also available before Einstein’s work (most famously, the Michelson-Morley
experiment). But his theory went much further. He radically reformulated the concepts
that we use to analyse force, energy, momentum, and so forth. In this sense, his new
theory was primarily a philosophical and conceptual achievement, rather than a new
experimental discovery of the kind traditionally regarded as the epitome of empirical
science.

He also attributed his universal ‘principle of relativity’ to the very nature of space and
time itself. With important contributions by Minkowski, this gave rise to the modern
view that physics is based on an inseparable combination of space and time, called
space-time. Minkowski treated this as a kind of ‘geometric’ entity, based on regarding
our Equation 1 as a ‘metric equation’ describing the geometric nature of space-time.
This view is called the ‘geometric explanation’ of relativity theory, and this approach
led Einstein even deeper into modern physics, when he applied this new conception to
the theory of gravity, and discovered a generalised theory of space-time.

The nature of this ‘geometric explanation’ of the connection between space, time, and
proper time is one of the most fascinating topics in the philosophy of physics. But it
involves the General Theory of Relativity, which goes beyond STR.

Time Supplement
This supplement answers a series of questions designed to reveal more about what
science requires of physical time, and to provide background information about other
topics discussed in the Time article.
Table of Contents

1. What are instants and durations?


2. What is an event?
3. What is a reference frame?
4. What is an inertial frame?
5. What is spacetime?
6. What is a Minkowski diagram?
7. What are the metric and the interval?
8. Does the theory of relativity imply time is partly space?
9. Is time the fourth dimension?
10. Is there more than one kind of physical time?
11. How is time relative to the observer?
12. What are the relativity and conventionality of simultaneity?
13. What is the difference between the past and the absolute past?
14. What is time dilation?
15. How does gravity affect time?
16. What happens to time near a black hole?
17. What is the solution to the twins paradox?
18. What is the solution to Zeno’s paradoxes?
19. How do time coordinates get assigned to points of spacetime?
20. How do dates get assigned to actual events?
21. What is essential to being a clock?
22. What is our standard clock?
23. Why are some standard clocks better than others?
24. What does it mean for a clock to be accurate?

1. What are instants and durations?


A duration is an amount of time. The duration of earth’s existence is about five billion
years; the duration of a flash of lightning is 0.0002 seconds. Years and seconds are not
durations; they are measures of durations. The second is the standard unit for the
measurement of time in the SI system (the International Systems of Units, that is, Le
Système International d’Unités). In informal conversation, an instant is a very short
duration. In physics, however, an instant is instantaneous; it is not a finite duration but
rather a “point” in time, or “a time.” The day begins at the instant called “midnight.”
It is an interesting question whether a finite duration of a real event is always a linear
continuum of instants, and, if so, how we know this.

2. What is an event?
In ordinary discourse, an event is a happening lasting a finite duration during which
some object changes its properties. For example, this morning’s event of buttering the
toast is the toast’s changing from unbuttered to buttered.

Without mentioning objects and properties, an event might instead be defined simply
as whatever is temporally before or after anything else. In physics, events are
considered to be more basic than objects and their properties. But if we do treat events
in terms of objects and properties, we might treat the buttering event as involving the
toast object having changed from not having the property of containing butter at a
certain time this morning at a certain location to its having the property of containing
butter at that location a few seconds later.

In ordinary discourse, an event has more than an infinitesimal duration, but in the
technical discourse of physics, all events are composed of point events, events with
zero duration and taking up zero volume of space (that is, being extensionless). Also, a
actual point event is considered by physicists to be a spacetime point’s having some
property other than those it has just by being a location in spacetime. By being so brief
and taking place in such a restricted volume, these point events are very idealized, but
they are very useful, and each location in spacetime is marked by an actual or possible
point event occurring there. For actual point events, the point event can be thought of
as the point’s having some property for an instant. Notice that no change is mentioned
here, nor is a physical object that has those properties. Point events are what all
objects and events are made of, and spacetime points are what have the properties.

A mathematical space is a collection of points, and the points might represent


anything, for example dollars. But the points of a real space such as spacetime are
actual and possible point events, and the points of a real space that is time are
instants.

These metaphysical assumptions of modern science are not part of common sense, the
shared background beliefs of most people. They also are not acceptable metaphysical
assumptions for many philosophers. In 1936, in order to avoid point events, Bertrand
Russell and A. N. Whitehead developed a theory of time based on the assumption that
all events in spacetime have a finite, non-zero duration. However, they had to assume
that any finite part of an event is an event, and this assumption is no closer to common
sense than the physicist’s assumption that all events are composed of point events.
The encyclopedia article on Zeno’s Paradoxes mentions that Michael Dummett and
Frank Arntzenius have continued in the 21st century to develop Russell’s and
Whitehead’s ideas about events having a finite, non-zero duration.

It is an open question in philosophy as to whether the passage of time is a feature of


the world to be explained by noting how events change, such as their changing from
being present to being past. Many philosophers believe it is improper to consider an
event to be something that can change.

For a more detailed discussion of what an event is, see the article Events.

3. What is a reference frame?


A reference frame for a space is a coordinate system, namely a standard point of view
or a perspective for making observations, measurements and judgments that assigns
unique values to each point of space.

Choosing a good reference frame can make a situation much easier to describe. If you
are trying to describe the motion of a car down a straight highway, you would not want
to choose a reference frame that is fixed to a spinning carousel. Instead, choose a
reference frame fixed to the highway. The motion of a planet is very complex as seen
from earth over many months. However, the motion is very simple in a frame of
reference at rest relative to the sun. Inertial frames are very special reference frames,
as we shall see below.

A reference frame is often specified by selecting a solid object that doesn’t change its
size and by saying that the reference frame is fixed to the object. We might select a
reference frame fixed to the Rock of Gibraltar. Another object is said to be at rest in the
reference frame if it remains at a constant distance in a fixed direction from the
reference body used to define the frame. For example, your house is at rest in a
reference frame fixed to the Rock of Gibraltar [not counting your house's vibrating
when a truck drives by, nor the house's speed due to plate tectonics]. When we say the
sun rose this morning, we are implicitly choosing a reference frame fixed to the earth’s
surface. The sun is not at rest in this reference frame.

The reference frame or coordinate system must specify locations, and this is normally
done by assigning numbers to points of space. In a three-dimensional space, the
analyst needs to specify four distinct points on the reference body, or four objects
mutually at rest somewhere in the frame. One point is the origin, and the other three
can be used to define three independent, perpendicular axes, the familiar x, y and z
directions, assuming a Cartesian or rectangular coordinate system were to be used.
Two point objects are at the same place if they have the same x-value, the same y-
value and the same z-value. To keep track of events, you will also need a time axis, a
“t” axis, and so you will expand your three-dimensional mathematical space to a four-
dimensional mathematical space. Two point events are simultaneous if they occur at
the same place and also at the same time. In this way, the analyst is placing a four-
dimensional coordinate system on the space and time. The coordinates could have
been letters instead of numbers, but numbers are the best choice because we want to
use them for measurement, not just for naming places.

The fact that physical spacetime has curvature implies that no single rigid (or
Cartesian) coordinate system is capable of covering the entire spacetime. To cover all
of spacetime in that case, we must make do with covering different regions of
spacetime with different coordinate patches that are “knitted together” where one
patch meets another. No single coordinate system can cover the surface of a sphere
without creating a singularity, but the sphere can be covered by patching together two
coordinate systems. Nevertheless if we can live with non-rigid curvilinear coordinates,
then any curved spacetime can be covered with a global four-dimensional coordinate
system in which every point being uniquely identified with a set of four numbers in a
continuous way.

That we use four numbers indicates the space is four-dimensional. And in creating
coordinate systems for spaces, the usual assumption is that we should supply n
independent numbers to specify a place in an n-dimensional space, where n is an
integer. This is usual but not required; instead we could exploit the idea that there are
space-filling curves which permit a single continuous curve to completely fill, and thus
coordinatize, a region of dimension higher than one, e.g., a plane. For this reason
(namely, that each point in n-dimensional space doesn’t always need n numbers to
name the point), the contemporary definition of “dimension” is rather exotic.

4. What is an inertial frame?


An inertial frame is either a non-accelerating frame, or, less generally, a reference
frame in which Newton’s laws of motion hold. Any spacetime obeying the laws of
Special Relativity can have an inertial frame across the whole universe. In General
Relativity any very small region of spacetime can have an inertial frame.

Suppose you’ve pre-selected your frame. How do you tell if you are in an inertial
frame? The answer is that you check that objects accelerate only when acted on by
forces. That is, you check that any object’s acceleration is zero if no net force acts on
the object. If no unbalanced external forces are acting on a moving object, then the
object moves in a straight line. It doesn’t curve; it coasts. And it travels equal distances
in equal amounts of time.

Any frame of reference moving at constant velocity relative to an inertial frame is also
an inertial frame. A reference frame spinning relative to an inertial frame is never an
inertial frame.

Einstein’s theory of special relativity is intended to apply only to inertial frames.


According to the theory, the speed of light in a vacuum is the same when observed
from any inertial frame of reference. Unlike the speed of a spaceship, the speed of light
in a vacuum isn’t affected by which inertial reference frame is used for the
measurement. If you have two relatively stationary, synchronized clocks in an inertial
frame, then they will read the same time, but if one moves relative to the other, then
they will get out of synchrony. This loss of synchrony due to relative motion is called
“time dilation.”

The presence of gravitation normally destroys any possibility of finding a perfect


inertial frame. So, in practice, when trying to use special relativity in a world containing
gravitation, inertial frames are distinguished not by being absolutely unaccelerated, but
rather by being unaccelerated relative to some suitably defined average of all the
matter in the universe. A reference frame in which star motion is ignored and the stars
are assumed to be at rest is approximately an inertial reference frame and is often
adequate for many purposes. This is the so-called inertial frame of the “fixed stars.”

5. What is spacetime?
Spacetime is where events are located, or, depending on your theory of spacetime, it’s
all possible events. Spacetime is a multi-dimensional space, one of whose dimensions is
time. It is often useful to suppose that there are four dimensions of spacetime. These
four include the time dimension of before-after and the three ordinary space
dimensions of, say, up-down, left-right, and forward-backward.

More technically, spacetime is the intended model of the general theory of relativity.
This requires it to be a differentiable space in which physical objects obey the
equations of motion of the theory. Minkowski space (that is, Minkowski spacetime) is
the model of special relativity. It’s a certain 4-dimensional real vector space. General
relativity theory requires that spacetime be locally like Minkowski spacetime.

Hermann Minkowski, in 1908, was the first person to say that spacetime is fundamental
and that space and time are just aspects of spacetime. Minkowski meant it is
fundamental in the sense that the spacetime interval between any two events is
intrinsic to spacetime and does not vary with the reference frame, unlike a spatial
distance or temporal duration.

Spacetime is a continuum in which we can define points and straight lines. However,
these points and lines do not satisfy the principles of Euclidean geometry when matter
is present. Einstein showed that the presence of matter affects geometry by warping
space and time. Einstein’s principal equation in his general theory of relativity implies
that the curvature of spacetime is directly proportional to the density of mass in the
spacetime. That is, Einstein says the structure of spacetime changes as matter moves
because the gravitational field from matter actually curves spacetime. Black holes are
a sign of radical curvature. The earth’s curving of spacetime is very slight but still
significant enough that it must be accounted for when synchronizing two Global
Positioning Satellites.

There have been serious attempts over the last few decades to construct theories of
physics in which spacetime is a product of more basic entities. The primary aim of
these new theories is to unify relativity with quantum theory. So far these theories
have not stood up to any empirical observations or experiments that could show them
to be superior to the presently accepted theories. So, for the present the concept of
spacetime remains fundamental.

The metaphysical question of whether spacetime is a substantial object or a


relationship among events, or neither, is considered in the discussion of the relational
theory of time.

6. What is a Minkowski diagram?


A spacetime diagram is a representation of the point-events of spacetime. In a
Minkowski spacetime diagram, a rectangular coordinate system is used, Einstein’s
Special Theory of Relativity holds, normally the time axis is vertical, one or two of the
spatial axes are suppressed, and an object’s inertial motion (coasting) produces events
in a straight line. Here is an example with space having just one dimension.

The above diagram shows Einstein standing still midway between the two places at
which there is a flash of light. The directed arrows represent the path of light rays from
the flash. In a Minkowski diagram, a physical object, such as an electron or a person’s
body, is not represented as occupying a point but as occupying a line containing all the
spacetime points at which it exists. The line, which usually isn’t straight, is called the
of the object. In the above diagram, Einstein’s worldline is a vertical line. If an object’s
worldline intersects or meets another object’s worldline, then the two objects have
collided. The units along the vertical time axis are customarily chosen to be the
product of time and the speed of light so that “worldlines” of light rays make a forty-
five degree angles with each axis. The set of all light speed world lines going through
an event defines the light cones of that event: the past light cone and the future light
cone.

Inertial motion produces a straight worldline, and accelerated motion produces a


curved worldline. If at some time Einstein were to jump on a train moving by at
constant speed, then his worldline would, from that time upward, tilt away from the
vertical and form some angle less than 45 degrees with the time axis. Events on the
same horizontal line of the Minkowski diagram are simultaneous in that reference
frame. A moving observer is added to this diagram to produce the diagram below in
the discussion about the relativity of simultaneity. In a coordinate system attached to
the Sun, the worldline of the Earth’s orbit would be a helix.

Not all spacetimes can be given Minkowski diagrams, but any spacetime satisfying
Einstein’s Special Theory of Relativity can. Minkowski diagrams are diagrams of a
Minkowski space, which is a spacetime satisfying the Special Theory of Relativity. This
theory falsely presupposes that physical processes, such as gravitational processes,
have no effect on the structure of spacetime. When attention needs to be given to the
real effect of these processes on the structure of spacetime, that is, when general
relativity needs to be used, then Minkowski diagrams become inappropriate for
spacetime. General relativity assumes that the geometry of spacetime is locally
Minkowskian but not globally. That is, spacetime is locally flat in the sense that in any
very small region one always finds spacetime to be 4-D Minkowskian (but not 4-D
Euclidean). Special relativity holds in any tiny region of spacetime that satisfies General
relativity, and so any tiny region can be fitted with an inertial reference frame. When
we say spacetime is “really curved” and not flat, we mean it really deviates from 4-D
Minkowskian geometry.

7. What are the metric and the interval?


A space is simply a collection of points. How far is it from one point to some different
point? The metric is the answer to this question. A metric on a space provides a
definition of distance (or length) by giving a function from each pair of nearby points to
a real number. In Euclidean space, the distance between two points is the length of the
straight line connecting them. This length is traditionally defined in terms of
coordinates using the Pythagorean Theorem.

Points are located by being assigned a coordinate. For doing science we want the
coordinate to be a real number, not, say, a letter of the alphabet. A coordinate for a
point in two-dimensional space requires two numbers; a coordinate for a point in n-
dimensional space requires n numbers. Time, being one-dimensional, requires a single
number. Time is considered a one-dimensional space mathematically, and the metric of
time is normally chosen to be the absolute value of the numerical difference between
the coordinates of the two points. For example, the duration between 5 AM and 8 AM is
three hours, assuming the times are for the same day.

In a 2-dimensional space, the metric is more complicated; the distance between the
point (x’,y’), with Cartesian coordinates x’ and y’, and the point (x,y), with coordinates
x and y, is defined to be the square root of (x’ – x) 2 + (y’ – y)2 when the space is flat,
that is, Euclidean. If the space is not flat, then a more sophisticated definition of the
metric is required. Note the application of the Pythagorean Theorem for Euclidean
space.

Our intuitive idea of what a distance is tells us that, however we define distance for a
space, we want it to have certain distance-like properties. For example, letting d(p,q)
stand for the distance between any two points p and q in the space, the following four
conditions must be satisfied:

1. d(p,p) = 0, and d(p,q) is greater than or equal to 0


2. If d(p,q) = 0, then p = q
3. d(p,q) = d(q,p)
4. d(p,q) + d(q,r) is greater than or equal to d(p,r)

Notice that there is no mention of the path the distance is taken across; all the
attention is on the point pairs themselves.

Do these conditions capture your idea of distance? If you were to check, you’d find that
the 2-D metric defined above, namely the square root of (x’ – x) 2 + (y’ – y)2, does
satisfy these four conditions. In 3-D Euclidean space, the metric that is defined to be
the square root of (x’ – x)2 + (y’ – y)2 + (z’ – z)2 works very well.
Consider the 4-D mathematical space that is used to represent the spacetime obeying
the laws of special relativity theory, namely Minkowski spacetime. What’s an
appropriate metric for this space? Well, if we were just interested in the space part of
this spacetime, then the above 3-D Euclidean metric is fine. But we’ve asked a delicate
question because the fourth dimension of this mathematical space is really a time
dimension and not a space dimension. Here is the so-called Lorentzian metric or
Minkowski metric for any pair of point events at (x’,y’,z’,t’) and (x,y,z,t) in Minkowskian
4-D spacetime:

Δs2 = – (x’ – x)2 – (y’ – y)2 – (z’ – z)2 + c2(t’ – t)2

Δs is called the interval of Minkowski spacetime. The interval corresponds to what


clocks measure between a pair of timelike events [that is, between a pair of events
separated enough in time that one could have had a causal effect on the other] and
what rulers measure between a pair of spacelike events. One other happy feature of
this metric is that the value of the interval is unaffected by changing to a new
reference frame provided the new reference frame is not accelerating relative to the
first. Changing from a first frame to a new, unaccelerated reference frame on the
spacetime will change the values of all the coordinates of the points of the spacetime,
but some relations between all pairs of points won’t be affected, namely the intervals
between pairs of points. Take any two observers who use a reference frame in which
they, themselves are fixed, and assume they are moving at constant velocity relative
to each other. Now consider some single event with a finite duration. The two observers
won’t agree on how long the event lasts, but they will always agree on the interval
between the beginning and end of the event.

The interval of spacetime is complicated because its square can be negative, unlike
with the space intervals we’ve discussed so far. If Δs2 is negative, the two points have
a space-like separation, meaning these events have a greater separation in space than
they do in time. If Δs2 is positive, then the two have a time-like separation, meaning
enough time has passed that one event could have had a causal effect on the other. If
Δs2 is zero, the two events might be identical, or they might have occurred millions of
miles apart. In ordinary space, if the space interval between two events is zero, then
the two events happened at the same time and place, but in spacetime, if the
spacetime interval between two events is zero, this means only that there could be a
light ray connecting them. All the events that have a zero spacetime interval from
some event e constitute e’s two light cones because they have the shape of cones
when represented in a Minkowski diagram, one cone for events in e’s future and one
cone for events in e’s past.

Because true metrics are always positive, the Lorentzian metric that we used above for
Minskowski spacetime is not a true metric, nor even a pseudometric; but it is
customary for physicists to refer to it loosely as a “metric” because Δs retains enough
other features of distance.

The interval in spacetime is an intrinsic feature of spacetime because it does not vary
with our choice of the coordinate system. The duration of an event is not intrinsic, nor
are the length or velocity of an object. In the Euclidean plane of 2-D space, length is
intrinsic. The metric determines the geometry of spacetime, which is always an intrinsic
feature of the spacetime. Adding space and time dependence to each term of the
Lorentzian metric produces the metric for general relativity.
8. Does the theory of relativity imply time is partly
space?
In 1908, Minkowski remarked that “Henceforth space by itself, and time by itself, are
doomed to fade away into mere shadows, and only a kind of union of the two will
preserve an independent reality.” Many people took this to mean that time is partly
space, and vice versa. C. D. Broad countered that the discovery of spacetime did not
break down the distinction between time and space but only their independence or
isolation. He argued that their lack of independence does not imply a lack of reality.
The Broad-Minkowski disagreement is still an issue in philosophy, but if Broad is
correct, then time is time; it’s not space at all.

Nevertheless, there is a deep sense in which time and space are “mixed up” or linked.
This is evident from the Lorentz transformations of special relativity that connect the
time t in one inertial frame with the time t’ in another frame that is moving in the x
direction at a constant speed v. In this equation, t’ is dependent upon the space
coordinate x and the speed. In this way, time is not independent of either space or
speed. It follows that the time between two events could be zero in one frame but not
zero in another. Each frame has its own way of splitting up spacetime into its space
part and its time part.

The reason why time is not partly space is that, within a single frame, time is distinct
from space. Time is not simply an arbitrary one-dimensional sub-space of spacetime; it
is a distinguished sub-space. That is, time is a distinguished dimension of spacetime,
not an arbitrary dimension. What being distinguished amounts to is that when you set
up a rectangular coordinate system on spacetime with an origin at, say, the event of
Mohammed’s birth, you may point the x-axis east or toward Mecca or away from the
center of Earth, but you may not point it forward in time–you may do that only with the
t-axis, the time axis.

9. Is time the fourth dimension?


Yes and no; it depends on what you are talking about. Time is the fourth dimension of
4-d spacetime, but time is not the fourth dimension of space, the space of places.

Mathematicians have a broader notion of the term “space” than the average person;
and in their sense a space need not consist of places, that is, geographical locations.
Not paying attention to the two meanings of the term “space” is the source of all the
confusion about whether time is the fourth dimension. The mathematical space used by
mathematical physicists to represent physical spacetime is four dimensional and in that
space, the space of places is a 3-d sub-space and time is another 1-d sub-space.
Minkowski was the first person to construct such a mathematical space, although in
1895 H. G. Wells treated time as a fourth dimension in his novel The time Machine.
Spacetime is represented mathematically by Minkowski as a space of events, not as a
space of ordinary geographical places.

In any coordinate system on spacetime, it takes at least four independent numbers to


determine a spacetime location. In any coordinate system on the space of places, it
takes at least three. That’s why spacetime is four dimensional but the space of places
is three dimensional. Actually this 19th century definition of dimensionality, which is
due to Bernhard Riemann, is not quite adequate because mathematicians have
subsequently discovered how to assign each point on the plane to a point on the line
without any two points on the plane being assigned to the same point on the line. The
idea comes from Georg Cantor. Because of this one-to-one correspondence, the points
on a plane could be specified with just one number. If so, then the line and plane must
have the same dimensions according to the Riemann definition. To avoid this problem
and to keep the plane being a 2-d object, the notion of dimensionality of a space has
been given a new, but rather complex, definition.

10. Is there more than one kind of physical time?


Every reference frame has its own physical time, but the question is intended in
another sense. At present, physicists measure time electromagnetically. They define
a standard atomic clock using periodic electromagnetic processes in atoms, then use
electromagnetic signals (light) to synchronize clocks that are far from the standard
clock. In doing this, are physicists measuring ‘”electromagnetic time” but not other
kinds of physical time?

In the 1930s, the physicists Arthur Milne and Paul Dirac worried about this question.
Independently, they suggested there may be very many time scales. For example,
there could be the time of atomic processes and light, which is measured best by
atomic clocks. There also could be the time of gravitation and large-scale physical
processes, which is measured best by the rotation of a pulsar (pulsating star). The two
physicists worried that the atomic clock and the astronomical clock might drift out of
synchrony after being initially synchronized, yet there would be no reasonable
explanation for why they don’t stay in synchrony. Ditto for clocks based on the
pendulum, on superconducting resonators, on the spread of electromagnetic radiation
through space, and on other physical principles. Just imagine the difficulty for physicists
if they had to work with electromagnetic time, gravitational time, nuclear time,
neutrino time, and so forth. Current physics, however, has found no reason to assume
there is more than one kind of time for physical processes.

In 1967, physicists did reject the astronomical standard for the atomic standard
because the deviation between known atomic and gravitation periodic processes could
be explained better assuming that the atomic processes were the more regular of the
two. Physicists had no reason to believe that a gravitational periodic process, that is
just as regular initially as the atomic process and that is not affected by friction or
impacts or other forces, would ever drift out of synchrony with the atomic process, yet
this is the possibility that worried Milne and Dirac.

11. How is time relative to the observer?


Physical time is not relative to any observer’s state of mind. Wishing time will pass
does not affect the rate at which the observed clock ticks. On the other hand, physical
time is relative to the observer’s reference system–in trivial ways and in a deep way
discovered by Albert Einstein.

In a trivial way, time is relative to the chosen coordinate system on the reference
frame, though not to the reference frame itself. For example, it depends on the units
chosen as when the duration of some event is 34 seconds if seconds are defined to be
a certain number of ticks of the standard clock, but is 24 seconds if seconds are
defined to be a different number of ticks of that standard clock. Similarly, the difference
between the Christian calendar and the Jewish calendar for the date of some event is
due to a different unit and origin. Also trivially, time depends on the coordinate system
when a change is made from Eastern Standard Time to Pacific Standard Time. These
dependencies are taken into account by scientists but usually never mentioned. For
example, if a pendulum’s approximately one-second swing is measured in a physics
laboratory during the autumn night when the society changes from Daylight Savings
Time back to Standard Time, the scientists do not note that one unusual swing of the
pendulum that evening took a negative fifty-nine minutes and fifty-nine seconds
instead of the usual one second.

Isn’t time relative to the observer’s coordinate system in the sense that in some
reference frames there could be fifty-nine seconds in a minute? No, due to scientific
convention, it is absolutely certain that there are sixty seconds in any minute in any
reference frame. How long an event lasts is relative to the reference frame used to
measure the time elapsed, but in any reference frame there are exactly sixty seconds
in a minute because this is true by definition. Similarly, you do not need to worry that in
some reference frame there might be two gallons in a quart.

In a deeper sense, time is relative, not just to the coordinate system, but to the
reference frame itself. That is Einstein’s principal original idea about time.

Einstein’s idea is that without reference to the frame, there is no fixed time interval
between two events, no ‘actual’ duration between them. Einstein illustrated his idea for
two observers, one on a moving train in the middle of the train, and a second observer
standing on the embankment next to the train tracks. If the observer sitting in the
middle of the rapidly moving train receives signals simultaneously from lightning
flashes at the front and back of the train, then in his reference frame the two lightning
strikes were simultaneous. But the strikes were not simultaneous in a frame fixed to an
observer on the ground. This outside observer will say that the flash from the back had
farther to travel because the observer on the train was moving away from the flash. If
one flash had farther to travel, then it must have left before the other one, assuming
that both flashes moved at the same speed. Therefore, the lightning struck the back of
the train before the lightning struck the front of the train in the reference frame fixed to
the tracks.

Let’s assume that a number of observers are moving with various constant speeds in
various directions. Consider the inertial frame of reference in which each observer is at
rest in his or her own frame. Which of these observers will agree on their time
measurements? Only observers with zero relative speed will agree. Observers with
different relative speeds will not, even if they agree on how to define the second and
agree on some event occurring at time zero (the origin of the time axis). If two
observers are moving relative to each other, but each makes judgments from a
reference frame fixed to themselves, then the assigned times to the event will disagree
more, the faster their relative speed. All observers will be observing the same objective
reality, the same event in the same spacetime, but their different frames of reference
will require disagreement about how spacetime divides up into its space part and its
time part.
This relativity of time to reference frame implies that there be no such thing as The
Past in the sense of a past independent of reference frame. This is because a past
event in one reference frame might not be past in another reference frame.

In some reference frame, was Adolf Hitler born before George Washington? No,
because the two events are causally connectible. That is, one event could in principle
have affected the other since light would have had time to travel from one to the other.
We can select a reference frame to reverse the usual earth-based order of two events
only if they are not causally connectible, that is, only if one event is in the absolute
elsewhere of the other. Despite the relativity of time to a reference frame, any two
observers in any two reference frames should agree about which of two causally
connectible events happened first.

12. What are the relativity and conventionality of


simultaneity?
If the universe obeys relativistic physics, then events that occur simultaneously with
respect to one reference frame will not occur simultaneously in another reference
frame that is moving with respect to the first frame. This is called the “relativity of
simultaneity.” It applies only to pairs of events in each other’s absolute elsewhere.

This Minkowski diagram represents Einstein sitting still in the reference frame while
Lorentz is traveling rapidly away from him and toward the source of flash 2. Because
Lorentz’s timeline is a straight line we can tell that he is moving at a constant speed.
The two flashes of light arrive at Einstein’s location simultaneously, creating spacetime
event B. However, Lorentz sees flash 2 before flash 1. That is, the event A of Lorentz
seeing flash 2 occurs before event C of Lorentz seeing flash 1. So, Einstein will readily
say the flashes are simultaneous, but Lorentz will have to do some computing to figure
out that the flashes are simultaneous in the frame because they won’t “look”
simultaneous. However, if we’d chosen a different reference frame from the one
above, one in which Lorentz is not moving but Einstein is, then Lorentz would be
correct to say flash 2 occurs before flash 1 in that new frame. So, whether the flashes
are or are not simultaneous depends on which reference frame is used in making the
judgment. It’s all relative.

This relativity of simultaneity is philosophically less controversial than the


conventionality of simultaneity. To appreciate the difference, consider what is involved
in making a determination regarding simultaneity. Given two events that happen
essentially at the same place, physicists assume they can tell by direct observation
whether the events happened simultaneously. If we don’t see one of them happening
first, then we say they happened simultaneously, and we assign them the same time
coordinate. The determination of simultaneity is more difficult if the two happen at
separate places, especially if they are very far apart. One way to measure
(operationally define) simultaneity at a distance is to say that two events are
simultaneous in a reference frame if unobstructed light signals from the two events
would reach us simultaneously when we are midway between the two places where
they occur, as judged in that frame. This is the operational definition of simultaneity
used by Einstein in his theory of relativity. Instead of using the midway method, we
could take the distant clock and send a signal home to our master clock, one already
synchronized with our standard clock; the master clock immediately sends a signal
back to the distant clock with the information about what time it was when the signal
arrived. We at the distant clock notice that the total travel time is t and that the master
clock’s signal says its time is, say, noon, so we immediately set our clock to be noon
plus half of t.

The “midway” method described above of operationally defining simultaneity in one


reference frame for two distant signals causally connected to us has a significant
presumption: that the light beams travel at the same speed regardless of direction.
Einstein, Reichenbach and Grünbaum have called this a reasonable “convention”
because any attempt to experimentally confirm it presupposes that we already know
how to determine simultaneity at a distance. This is the conventionality, rather than
relativity, of simultaneity. To pursue the point, suppose the two original events are in
each other’s absolute elsewhere; they couldn’t have affected each other. Einstein
noticed that there is no physical basis for judging the simultaneity or lack of
simultaneity between these two events, and for that reason said we rely on a
convention when we define distant simultaneity as we do. Hillary Putnam, Michael
Friedman, and Graham Nerlich object to calling it a convention–on the grounds that to
make any other assumption about light’s speed would unnecessarily complicate our
description of nature, and we often make choices about how nature is on the basis of
simplification of our description. They would say there is less conventionality in the
choice than Einstein supposed.

The “midway” method isn’t the only way to define simultaneity. Consider a second
method, the “mirror reflection” method. Select an earth-based frame of reference, and
send a flash of light from earth to Mars where it hits a mirror and is reflected back to its
source. The flash occurred at 12:00, let’s say, and its reflection arrived back on earth
20 minutes later. The light traveled the same empty, undisturbed path coming and
going. At what time did the light flash hit the mirror? The answer involves the so-called
conventionality of simultaneity. All physicists agree one should say the reflection event
occurred at 12:10. The controversial philosophical question is whether this is really a
convention. Einstein pointed out that there would be no inconsistency in our saying
that it hit the mirror at 12:17, provided we live with the awkward consequence that
light was relatively slow getting to the mirror, but then traveled back to earth at a
faster speed. If we picked the impact time to be 12:05, we’d have to live with the fact
that light traveled slower coming back. There is a physical basis for not picking the
impact time to be less than noon nor later than 12:20, because doing so would violate
the physical principle that causes precede their effects. One requirement we place on
the concept of simultaneity is that distant events which are simultaneous could not be
in causal contact with each other. We can satisfy that requirement for any choice of
impact time from 12:00 to 12:20.

13. What is the difference between the past and


the absolute past?

The events in your absolute past are those that could have directly or indirectly
affected you, the observer, now. These absolutely past events are the events in or on
the backward light cone of your present event, your here-and-now. The backward light
cone of event Q is the imaginary cone-shaped surface of spacetime points formed by
the paths of all light rays reaching Q from the past. An event’s being in another event’s
absolute past is a feature of spacetime itself because the event is in the point’s past in
all possible reference frames. The feature is frame-independent. For any event in your
absolute past, every observer in the universe (who isn’t making an error) will agree the
event happened in your past. Not so for events that are in your past but not in your
absolute past. Past events not in your absolute past will be in what Eddington called
your “absolute elsewhere” and these past events will be in your present as judged by
some other reference frames. The absolute elsewhere is the region of spacetime
containing events that are not causally connectible to your here-and-now. Your
absolute elsewhere is the region of spacetime that is neither in nor on either your
forward or backward light cones. No event here now, can affect any event in your
absolute elsewhere; and no event in your absolute elsewhere can affect you here and
now. A spacetime point’s absolute future is all the future events outside the point’s
absolute elsewhere.

A single point’s absolute elsewhere, absolute future, and absolute past partition all of
spacetime beyond the point into three disjoint regions. If point A is in point B’s
absolute elsewhere, the two events are said to be “spacelike related.” If the two are in
each other’s forward or backward light cones they are said to be “timelike related” or
“causally connectible.”

14. What is time dilation?


According to special relativity, two properly functioning clocks next to each will stay
synchronized. Even if they were to be far away from each other, they’d stay
synchronized. But if one clock moves away from the other, the moving clock will tick
slower than the stationary clock, as measured in the inertial reference frame of the
stationary clock. This slowing due to motion is called “time dilation.” If you move at
99% of the speed of light, then your time slows by a factor of 7 relative to stationary
clocks. In addition, you are 7 times thinner than when you are stationary, and you are 7
times heavier. If you move at 99.9%, then you slow by a factor of 22.

Time dilation is about clocks in different frames disagreeing with each other. Suppose
your twin’s spaceship travels to and from a star one light year away. It takes light from
your Earth-based flashlight two years to go there and back. But if the spaceship is fast,
your twin can make the trip in less than two years, according to his own clock. Does he
travel the distance in less time than it takes light to travel that distance? No, according
to yourclock he takes more than two years, and so is slower than light.

We sometimes speak of time dilation by saying time itself is “slower,” but time isn’t
going slower in any absolute sense, only relative to some other frame of reference.
Does time have a rate? Well, time in a reference frame has no rate in that frame, but
time in a reference frame can have a rate as measured in a different frame, such as in
a frame is moving relative to the first frame.

Time dilation is not an illusion of perception; and it’s not a matter of the second having
different definitions in different reference frames. Also, it’s not a Doppler effect. Time
dilation isn’t affected by direction of motion. The Doppler effect is affected by direction
of motion, which we detect in the difference between a blue shift and a red shift.

Time dilation due to difference in constant speeds is described by Einstein’s special


theory of relativity. The general theory of relativity describes a second kind of time
dilation, one due to different accelerations and different gravitational influences. For
more on general relativistic dilation, see the discussion of gravity and black holes.

Newton’s physics describes duration as an absolute property, implying it is not relative


to the reference frame. However, in Newton’s physics the speed of light is relative to
the frame. Einstein’s special theory of relativity reverses both of these aspects of time.
For inertial frames, it implies the speed of light is not relative to the frame, but
duration is relative to the frame. In general relativity, however, the speed of light can
vary within one reference frame if matter and energy are present.

Time dilation due to motion is relative in the sense that if your spaceship moves past
mine so fast that I measure your clock to be running at half speed, then you will
measure my clock to be running at half speed also, provided both of us are in inertial
frames. If one of us is affected by a gravitational field or undergoes acceleration, then
that person isn’t in an inertial frame and the results are different.

Both types of time dilation play a significant role in time-sensitive satellite navigation
systems such as the Global Positioning System. The atomic clocks on the satellites
must be programmed to compensate for the relativistic dilation effects of both gravity
and motion.

15. How does gravity affect time?


Einstein’s general theory of relativity (1915) is a generalization of his special theory of
special relativity (1905). It is not restricted to inertial frames, and it encompasses a
broader range of phenomena, namely gravity and accelerated motions. According to
general relativity, gravitational differences affect time by dilating it. Observers in a less
intense gravitational potential find that clocks in a more intense gravitational potential
run slow relative to their own clocks. People live longer in basements than in attics, all
other things being equal. Basement flashlights will be shifted toward the red end of the
visible spectrum compared to the flashlights in attics. This effect is known as the
gravitational red shift. Even the speed of light is slower in the presence of higher
gravity.

Informally one speaks of gravity bending light rays around massive objects, but more
accurately it is the space that bends, and as a consequence the light is bent, too. The
light simply follows the shortest path through spacetime, and when space curves the
shortest paths are no longer Euclidean straight lines.

16. What happens to time near a black hole?


A black hole is a volume of very high gravitational field or severe warp in the spacetime
continuum. Astrophysicists believe black holes are commonly formed by the inward
collapse of stars that have burned out. The center of a spherical black hole is infinitely
dense according to relativity theory, but some theories of quantum gravity imply that
the density cannot reach infinity. It is surrounded by an event horizon, a concentric
sphere marking the point of no return. Anything getting that close could never escape
the inward pull, even if it had an unlimited fuel supply and could travel at near the
speed of light. Anything crossing the event horizon from the outside would quickly
crash into the center of the black hole and be crushed to a point, according to relativity
theory. The singularity is the point of infinite density in the black hole. The first black
hole solution to Einstein’s equations of general relativity were discovered by
Schwarzschild in 1916. Because even light itself could not escape from inside a black
hole, John Wheeler chose the name “black hole.”
In relativity theory, the proper time between two events along a worldline is the time
that would be shown on a clock whose path in spacetime is that worldline between the
events. The proper time is not the same as thecoordinate time. Coordinate time is time
along the worldline of an ideal clock at the origin of the coordinate system. The
coordinate time between the two events is the time separation of the events given by
an observer at rest in the frame. Proper time is independent of coordinate time,
although the usual convention is to measure both times in the same units, namely
seconds. As judged by a clock on earth in an earth-based frame of reference, an
astronaut flying into a distant black hole will take an infinite coordinate time to reach
the event horizon of the black hole. That is, if we could see the astronaut’s clock, the
clock would appear to us to slow to a halt. But as judged by the astronaut, it will take
only a few microseconds of the astronaut’s proper time to pass through the event
horizon and crash into the center of the black hole.

If you, the person falling toward the event horizon, were to escape the pull towards the
black hole and return home, you’d discover that you were younger than your earth-
bound twin and that your initially synchronized clocks showed that yours had fallen
behind. It is in this sense that you’ve experienced a time warp, a warp in the time
component of spacetime. According to Stanford physicist Leonard Susskind, there is “a
very common misconception, namely, that because an outside observer sees an
infalling observer slow down, that the infalling observer sees the outsider speed up.
This is simply not so. The infalling observer looks back and sees nothing unusual.”

17. What is the solution to the twins paradox?


This paradox, also called the clock paradox and the twin paradox, is an argument
about time dilation that uses the theory of relativity to produce a contradiction. The
argument considers two twins at rest with their clocks synchronized. One twin climbs
into a spaceship and flies far away at a constant speed, then reverses course and flies
back at the same speed. When they reunite, will the twins still be the same age? No.
Relativity theory implies that the twin on the spaceship will return and be younger than
the Earth-based twin. The elapsed proper time of the twin who returns is less than the
elapsed proper time of the Earth-based twin. However, it’s all relative, isn’t it? That is,
we could have considered the spaceship to be stationary. Wouldn’t relativity theory
then imply that the Earth-based twin would race off (along with the Earth), then return
and be the younger of the two? If so, we have a contradiction: when the twins reunite,
each will be younger than the other.

Einstein worried about the paradox [Einstein, Naturwissenschaften, 6, 697 (1918)], and
Herbert Dingle famously argued in the 1960’s that the paradox reveals an
inconsistency in special relativity. Almost all philosophers and scientists now agree
that it is not a true paradox, in the sense of revealing a logical inconsistency within
relativity theory, but is merely a complex puzzle that can be adequately solved within
relativity theory. The twin who feels the acceleration is the twin who becomes the
younger twin, but the acceleration upon starting and reversing course is not what
causes this difference in aging, and it is not essential to the paradox. The key idea is
that there is an asymmetry in the two spacetime intervals taken by the twins between
the goodbye event and the reunion. Sitting still on Earth is a way of maximizing the
time between the events; flying fast in a spaceship is a way of minimizing the time
between the events. The reasoning in the paradox makes the mistake of supposing
that the twin on Earth could somehow have the shorter interval.

To explain that last point, let’s for simplicity’s sake assume the twin on Earth is fixed in
an inertial frame. The way out of the paradox is to notice that the argument has two
halves. The first half describes the twin in the spaceship flying away, then turning
around and flying back to the Earth-based twin who remains fixed in an inertial frame
during the flight. The second half describes the Earth-based twin as flying away from
the spaceship and then returning to the spaceship, while the spaceship remains
stationary in an inertial frame during the flight. The problem is that in the second half
of the paradox it was a mistake to suppose the spaceship “remains stationary in an
inertial frame.” The spaceship’s frame can not be inertial. Also, because of the
spaceship’s changing velocity by turning around, the twin on the spaceship has a
shorter world-line than the Earth-based twin and takes less time than the Earth-based
twin. The assumption is that the stars are not moving in tandem with the spaceship but
are generally stationary relative to the Earth. Without this crucial, but usually implicit,
assumption, one couldn’t decide which twin was or was not in an inertial frame; if two
twins are out in space alone with no stars (or other matter-energy), then when they
meet again they will be the same age.

The production of the paradox depends on using a heuristic principle that the
description of the world is equally valid from the point of view of any observer. That
heuristic principle is embedded in the above remark, “…it’s all relative, isn’t it?” The
application of the principle makes the assumption that the two halves of the analysis
are working with two equivalent descriptions of the same process. However, the two
descriptions are not equivalent because, if there is an inertial frame in the first half of
the argument, then there is no available inertial frame for the second half.

The analyst is always free to make the choice of a non-inertial frame in which the
spaceship is considered to be stationary. This would complicate the analysis, and
require general relativity instead of special relativity, but the result would be the same,
namely no contradiction. So, it can’t be shown that the Earth-based twin is the
younger, and the paradox is merely a puzzle that has a solution.

To dig more deeply into the cause of error in the reasoning, notice that the production
of the paradox depends upon a careless use of the heuristic principle that the
description of the world is equally valid from the point of view of any observer. This
principle is misinterpreted in the twins paradox. What is always correct in relativity
theory and what underlies the heuristic principle is the symmetry principle: the
invariance of the laws under Lorentz transformations. These transformation equations
give the relations between the coordinates of a single event [such as our spaceship’s
flight away from the Earth-based twin] as measured by observers in two
different inertial reference frames in motion relative to one another. But there aren’t
two inertial frames to use in the case of the twins paradox, so the symmetry principle is
correct, but the heuristic principle is not applicable. The argument of the twins paradox
applies the heuristic principle anyway, and draws an incorrect conclusion that there is a
contradiction.

What causes one twin to age differently? The best answer to this question is to re-
examine the question itself. It was remarked above that the easiest way to see the
dissimilarity in the two halves of the argument. Failure to notice the asymmetry in the
two halves is the cause of the error in the paradox, but it’s not the cause of the age
difference in the twins. Their age difference isn’t caused by anything, just as light’s
going at the speed of light instead of at some other speed isn’t caused by something
but is just the way nature behaves, at least insofar as the theory of relativity is
concerned.

18. What is the solution to Zeno’s paradoxes?


See the article “Zeno’s Paradoxes” elsewhere in this encyclopedia.

19. How do time coordinates get assigned to points


of spacetime?
A space is a collection of points. In a space that is supposed to be time, these points
are the instants. Our question is how we assign time numbers to these points. Before
discussing time coordinates specifically, let’s consider what is meant by assigning
coordinates to a mathematical space, one that might represent either physical space,
or physical time, or spacetime, or something else. In a one-dimensional space, such as
a curved line, we assign unique coordinate numbers to points along the line and we
make sure that no point fails to have a coordinate. For a 2-dimensional space, we
assign pairs of numbers to points. For a 3-d space, we assign triples of numbers. If we
assign letters instead of numbers, we can’t use the tools of mathematics to describe
the space. But we can’t assign any coordinate numbers we please. There are
restrictions. For example, if the space has a certain geometry, then we have to assign
numbers that reflect this geometry. Here is the fundamental method of analytic
geometry:

Consider a space as a class of fundamental entities: points. The class of points has
“structure” imposed upon it, constituting it a geometry–say the full structure of space
as described by Euclidean geometry. [By assigning coordinates] we associate another
class of entities with the class of points, for example a class of ordered n-tuples of real
numbers [for a n-dimensional space], and by means of this “mapping” associate
structural features of the space described by the geometry with structural features
generated by the relations that may hold among the new class of entities–say
functional relations among the reals. We can then study the geometry by studying,
instead, the structure of the new associated system [of coordinates]. (Sklar, 1976, p.
28)

The goal in assigning coordinates to a space is to create a reference system for the
space. A reference system is a reference frame plus either a coordinate system or
an atlas of coordinate systems placed by the analyst upon the space to uniquely name
the points. These names or coordinates are frame dependent in that a point can get
new coordinates when the reference frame is changed. For 4-d spacetime obeying
special relativity and its Lorentzian geometry, a coordinate system is a grid of
smooth timelike and spacelike curves on the spacetime that assigns to each point three
space coordinate numbers and one time coordinate number. No two distinct points can
have the same set of four coordinate numbers. Inertial frames can have global
coordinate systems, but in general we have to make due with atlases. If we are working
with general relativity where spacetime can curve and we cannot assume inertial
frames, then the best we can do is to assign a coordinate system to a small region of
spacetime where the laws of special relativity hold to a good approximation. General
relativity requires special relativity to hold locally, and thus for spacetime to be
Euclidean locally. That means that locally the 4-d spacetime is correctly described by 4-
d Euclidean solid geometry. Consider two coordinate systems on adjacent regions. For
the adjacent regions we make sure that the ‘edges’ of the two coordinate systems
match up in the sense that each point near the intersection of the two coordinate
systems gets a unique set of four coordinates and that nearby points get nearby
coordinate numbers. The result is an “atlas” on spacetime.

For small regions of spacetime, we create a coordinate system by choosing a style of


grid, say rectangular coordinates, fixing a point as being the origin, selecting one
timelike and three spacelike lines to be the axes, and defining a unit of distance for
each dimension. We cannot use letters for coordinates. The alphabet’s structure is too
simple. Integers won’t do either; but real numbers are adequate to the task. The
definition of “coordinate system” requires us to assign our real numbers in such a way
that numerical betweenness among the coordinate numbers reflects the betweenness
relation among points. For example, if we assign numbers 17, pi, and 101.3 to instants,
then every interval of time that contains the pi instant and the 101.3 instant had better
contain the 17 instant. When this feature holds, the coordinate assignment is said to be
monotonic. There is no way to select one point of spacetime and call it the origin of the
coordinate system except by reference to actual events. In practice, we make the
origin be the location of a special event, and one popular choice is the birth of Jesus.
Negative time coordinates are assigned to events occuring before the birth of Jesus.

The choice of the unit presupposes we have defined what “distance” means.
The metric for a space specifies what is meant by distance in that space. The natural
metric between any two points in a one-dimensional space, such as the time sub-space
of our spacetime, is the numerical difference between the coordinates of the two
points. Using this metric, the duration between the 11:00 instant and the 11:05 instant
is five minutes. The metric for spacetime defines the spacetime interval between two
spacetime locations, and it is more complicated than the metric for time alone. The
spacetime interval between any two events is invariant or unchanged by a change to
any other reference frame, although the spatial distances and durations do vary. More
accurately, in the general theory, the infinitesimal spacetime interval between two
neighboring points is invariant. The units of the spacetime interval are seconds
squared.

In this discussion, there is no need to worry about the distinction between change in
metric and change in coordinates. For a space that is topologically equivalent to the
real line and for metrics that are consistent with that topology, each coordinate system
determines a metric and each metric determines a coordinate system. More precisely,
once you decide on a positive direction in the one-dimensional space and a zero-point
for the coordinates, then the possible coordinate systems and the possible metrics are
in one-to-one correspondence.

There are still other restrictions on the assignments of coordinate numbers. The
restriction that we called the “conventionality of simultaneity” fixes what time-slices of
spacetime can be counted as collections of simultaneous events. An even more
complicated restriction is that coordinate assignments satisfy the demands of general
relativity. The metric of spacetime in general relativity is not global but varies from
place to place due to the presence of matter and gravitation. Spacetime cannot be
given its coordinate numbers without our knowing the distribution of matter and
energy.

The features that a space has without its points being assigned any coordinates
whatsoever are its topological features. These are its dimensionality, whether it goes
on forever or has a boundary, how many points there are, and so forth.

20. How do dates get assigned to actual events?


Our purpose in choosing a coordinate system or atlas to assign real numbers to all
spacetime points is to express relationships among actual and possible events. The
relationships we are interested in are order relationships (Did this event occur between
those two?) and magnitude relationships (How long after A did B occur?). The date of a
(point) event is the time coordinate number of the spacetime location where the event
occurs. We expect all these assignments of dates to events to satisfy the requirement
that event A happens before event B iff t(A) < t(B), where t(A) is the time coordinate of
A. The assignments of dates to events also must satisfy the demands of our physical
theories, and in this case we face serious problems involving inconsistency as when a
geologist gives one date for the birth of earth and an astronomer gives a different date.

It is a big step from assigning numbers to points to assigning them to real events. Here
are some of the questions that need answers. How do we determine whether a nearby
event and a distant event occurred simultaneously? Assuming we want the second to
be the standard unit for measuring the time interval between two events, how do we
operationally define the second so we can measure whether one event occurred
exactly one second later than another event? How do we know whether the clock we
have is accurate? Less fundamentally, attention must also be paid to the dependency
of dates due to shifting from Standard Time to Daylight Savings Time, to crossing the
International Date Line, and to switching from the Julian to the Gregorian Calendar.

Let’s design a coordinate system. Suppose we have already assigned a date of zero to
the event that we choose to be at the origin of our coordinate system. To assign dates
to other events, we first must define a standard clock and declare that the time
intervals between any two consecutive ticks of that clock are the same. The second,
our conventional unit of time measurement, will be defined to be so many ticks of
the standard clock. We then synchronize other clocks with the standard clock so the
clocks show equal readings at the same time. The time at which a point event occurs is
the number reading on the clock at rest there. If there is no clock there, the assignment
process is more complicated.

We want to use clocks to assign a time even to distant events, not just to events in the
immediate vicinity of the clock. To do this correctly requires some appreciation of
Einstein’s theory of relativity. A major difficulty is that two nearby synchronized clocks,
namely clocks that have been calibrated and set to show the same time when they are
next to each other, will not in general stay synchronized if one is transported
somewhere else. If they undergo the same motions and gravitational influences, they
will stay synchronized; otherwise, they won’t. For more on how to assign dates to
distant events, see the discussion of the relativity and conventionality of simultaneity.
As a practical matter, dates are assigned to events in a wide variety of ways. The date
of the birth of the Sun is measured very differently from dates assigned to two
successive crests of a light wave. For example, there are lasers whose successive
crests of visible light waves pass by a given location every 10 to the minus 15 seconds.
This short time isn’t measured with a stopwatch. It is computed from measurements of
the light’s wavelength. We rely on electromagnetic theory for the equation connecting
the periodic time of the wave to its wavelength and speed. Dates for other kinds of
events also are often computed rather than directly measured with a clock.

21. What is essential to being a clock?


Clocks record numerical information about time. They measure the quantity of time,
the duration. Every clock has two functions. One function is to generate a sequence of
events of hopefully the same durations. Periodic processes provide these events. In a
wall clock, the events are pairs of successive ticks. In a pendulum clock, the events are
swings (that is, oscillations) of the pendulum. The second function is to count these
events, thereby providing a measurement of their durations in seconds and minutes
and hours and years. This counting can be especially difficult if the ticks are occurring a
trillion times a second. However, it is an arbitrary convention that we design clocks to
count up to higher numbers rather than down to lower numbers as time goes on.

One principal goal in clock-building is to make the clock’s basic durations be congruent.
That is, the duration between any two adjacent ticks should be the same. When this
goal is achieved, the clock is said to be uniform or regular.

A second goal is for the time measurements of the clock to agree with those of
the standard clock. When this happens, the clock is said to be properly calibrated
or accurate or synchronized with the standard clock. To calibrate a clock, that is, to
synchronize it with the standard clock, we want our clock to show that it is time t just
when the standard clock shows that it is time t, for all t.

A clock isn’t really measuring the time in a reference frame other than one fixed to the
clock. In other words, a clock primarily measures the elapsed proper time between
events that occur along its own worldline. Technically, a clock is a device that
measures the spacetime interval along its own worldline. If the clock is at rest in an
inertial frame, then it measures the “coordinate time.” If the spacetime has no inertial
frame then it can’t have a coordinate time. Because clocks are intended to be used to
measure events external to themselves, a third goal in clock building is to ensure there
is no difficulty in telling which clock tick is simultaneous with which events occuring
away from the clock. For example, we might want to determine when the sun comes up
in the morning at some particular place where we and our clock are located. For some
clocks, the sound made by the ticking helps us make this determination. For other
clocks, the determination is made by our seeing the sun rise just when we see the
digital clock face show a specific time of day. More accuracy in the determination
requires less reliance on human judgment.

In our discussion so far, we have assumed that the clock is very small, that it can count
any part of a second and that it can count high enough to be a calendar. This isn’t
always a good assumption with a real clock. Despite the practical problems, there is the
problem of there being a physical limit to the shortest duration measurable by a given
clock because no clock can measure time more accurately than the time it takes light
to travel between the components of that clock, the components in the part that
generates the sequence of regular ticks.

22. What is our standard clock?


By current convention [in 1964 by ratification by the General Conference of Weights
and Measures for the International System of Units, which replaced what was called the
"metric system"], the standard clock is the clock we agree to use for defining the
standard second. The current standard second is defined to be the duration of
9,192,631,770 periods (cycles, oscillations, vibrations) of a certain kind of microwave
radiation in the standard clock. More specifically, the second is defined to be the
duration of 9,192,631,770 periods of the microwave radiation required to produce the
maximum fluorescence of cesium 133 atoms (that is, their radiating a specific color of
light) as the atoms make a transition between two specific hyperfine energy levels of
the ground state of the atoms. This is the internationally agreed upon unit for atomic
time [the T.A.I. system]. The old astronomical system [Universal Time 1] defined a
second to be 1/86,400 of an Earth day.

For atomic time, the atoms of cesium with a uniform energy are sent through a
chamber that is being irradiated with these microwaves. The frequency of these
microwaves is tuned until the maximum number of cesium atoms flip from one energy
to the other, showing that the microwave radiation frequency is now precisely tuned to
be 9,192,631,770 vibrations per second. Because this frequency for maximum
fluorescence is so stable from one experiment to the next, the vibration number is
accurate to so many significant digits. The National Institute of Standards and
Technology’s F-1 atomic fountain clock, which was adopted in late 1999 as the primary
time standard of the United States, is so accurate that it drifts by less than one second
every 20 million years. We know there is this drift because it is implied by the laws of
physics, not because we have a better clock from which to make the judgment.

The standard clock is used to fix the units of all lengths. The unit of length depends on
the unit of time. The meter depends on the second. It does not follow from this, though,
that time is more basic than space. All that follows is that time measurement is more
basic than space measurement. And this has to do with convention and with the fact
that current science is capable of measuring time more precisely than space.

Thanks to the regularity of light propagation in a vacuum, the meter is defined in terms
of the second. The meter is defined in terms of the pre-defined second as being the
distance light travels in exactly 0.000000003335640952 seconds or 1/299,792,458
seconds. That number is picked so that the new meter will be nearly the same distance
as the old meter, which was the distance between two marks on a platinum bar that
was kept in the Paris Observatory. Why is the meter defined in terms of the second,
instead of having the second defined in terms of the meter as, say, how long it takes
light to travel a certain distance? The answer is that distance can’t be measured as
accurately as time. Time can be more accurately measured than distance, voltage,
temperature, mass, or anything else.

These standard definitions of the second and the meter amount to defining or fixing the
speed of light in a vacuum in all inertial frames. The speed is exactly one meter per
0.000000003335640952 seconds or 299,792,458 meters per second, or approximately
186,282 miles per second or about a foot per nanosecond. There can no longer be any
direct measurement to see if that is how fast light REALLY moves in an inertial frame; it
is simply defined to be moving that fast. Any measurement that produced a different
value for the speed of light would be presumed initially to have an error in, say, its
measurements of lengths and durations, or in its assumptions about the influence of
gravitation and acceleration, or in its assumption that the light was moving in a
vacuum. This initial presumption comes from a deep reliance by scientists on Einstein’s
theory of relativity. However, if it were eventually decided by the community of
scientists that the theory of relativity is incorrect and that the speed of light shouldn’t
have been fixed as it was, then the scientists would call for a new world convention to
re-define the second. Some physicists believe that a better system of units would first
define the speed of light, then define the second, and then make the meter be a
computed consequence of these.

Although a microwave atomic clock is currently used for our standard unit of time, it is
expected that in the first quarter of the 21st century, physicists will agree to use an
optical atomic clock, and then the definition of the second will be changed to refer to
an optical frequency, rather than to a microwave frequency.

23. Why are some standard clocks better than


others?
We choose as our standard clock our best clock, the one with the least drift, the one
with the most regularity in its period. Other clocks ideally are calibrated by being
synchronized to this standard clock.

In about 1700, scientists discovered that their best watches and clocks showed that the
time from one day to the next, as determined by sunrises, varied throughout the year.
Therefore, they preferred to define durations in terms of the mean or average day
throughout the year. Before the 1950s, the standard clock was defined astronomically
in terms of the mean rotation of the earth upon its axis [solar time]. For a short period
in the 1950s and 1960s, it was defined in terms of the revolution of the earth about the
Sun [ephemeris time]. The second was defined to be 1/86,400 of the mean solar day,
the average throughout the year of the rotational period of the earth with respect to
the Sun. Now we’ve found a better standard clock, a certain kind of atomic clock
[atomic time]. All atomic clocks measure time in terms of the natural resonant
frequencies of various atoms and molecules. The periodic behavior of a super-cooled
cesium atomic clock is the best practical standard clock we have so far discovered.
[The dates of adoption of the standards was left vague in the previous sentences
because different international organizations adopted different standards in different
years.]

The principal theoretical goal in selecting a standard clock is to find a periodic (cyclic)
process that, if adopted as our standard, makes the resulting system of physical laws
simpler and more useful. Choosing the atomic clock as standard is much better for
this purpose than choosing the periodic dripping of water from our goat skin bag or the
period of a special pendulum or even the periodic revolution of the earth about the
Sun.
When we choose a standard clock we are making a choice about how to compare two
durations in order to decide whether they are of equal duration. Is this choice somehow
forced upon us? To what extent is this choice conventional? Philosophers dispute the
extent to which the choice of metric is conventional rather than forced by nature.
Taking the conventional side, Adolf Grünbaum argues that time is metrically
amorphous. It has no intrinsic metric in the sense of its structure determining the
measure of durations. Instead, we analysts establish durations between instants by the
way we assign coordinates to instants. If we were to say the instant at which Jesus was
born and the instant at which Abraham Lincoln was assassinated occurred only 24
seconds apart, whereas the duration between Lincoln’s assassination and his burial is
24 billion seconds, then we can’t be mistaken. It’s up to us to say what is correct when
we first create our conventions about measuring duration. We can consistently assign
any numerical time coordinates we wish, subject only to the condition that the
assignment properly reflect the betweenness relations of the events that occur at those
instants. That is, if event J (birth of Jesus) occurs before event L (Lincoln’s
assassination) and this in turn occurs before event B (burial of Lincoln), then the time
assigned to J must be numerically less than the time assigned to L, and both must be
less than the time assigned to B. t(J) < t(L) < t(B). A simple requirement. It is other
requirements that lead us to reject the above convention about 24 seconds and 24
billion seconds as unhelpful. What requirements? We’ve found that, for doing science,
certain processes are more “regular” than others. Pendulum swings are more regular
than repeated barks of a dog. Periodic appearances of the sun overhead are more
regular than rainstorms. Why are they? It’s because there are many periodic processes
in nature that have a special relationship to each other; their periods are very nearly
constant multiples of each other, and this constant stays the same over a long time.
For example, the period of the rotation of the Earth is a fairly constant multiple of the
period of the revolution of the Earth around the Sun, and both these periods are a
constant multiple of the periods of swinging pendulums. The class of these periodic
processes is very large, so the world will be easier to describe if we choose our
standard clock from one of these periodic processes. If we were to choose the standard
to be the period of our own pulse, then we’d find that all those other processes would
speed up when we are excited and slow down when we are not, and we’d find that it
would be more difficult to find simple laws of nature. A good convention for what is
regular will make it easier for scientists to find simple laws of nature and to explain
what causes other events to be irregular. It is the search for regularity and simplicity
that leads us to adopt the conventions for numerical time coordinate assignments that
we do. No, says the objectivist, the success of the atomic clock over these other clocks
we might have chosen as our standard clock shows that we picked the correct clock. An
objectivist believes that whether two intervals of time are really equivalent is an
intrinsic feature of nature, and choosing an atomic clock instead of the earth’s
revolutions about the sun as the standard clock isn’t any more conventional than is
choosing to say the Earth is round rather than flat.

A practical goal in selecting a standard clock is to find a clock that is relatively


insulated from environmental impact such as comet impacts, stray electric fields or the
presence of dust. If not insulation, then compensation. That is, if there is some
theoretically predictable effect upon the standard clock, then the clock can be regularly
adjusted to take account of the effect. Sensors, such as a thermometer or whatever,
will sense the local conditions that affect the clock, and their readings can be used to
apply a suitable correction in order to compensate for the effect of those conditions.
Why is choosing the cesium atomic clock as our standard better than choosing an
astronomical process such as the mean yearly motion of the earth around the Sun? The
brief answer is that the earth’s rate of spin varies. The ocean’s tides, the sloshing of
earth’s molten core, and other things, are affecting the rotation of the earth, but not
affecting the atomic clock. If we said that by definition the earth doesn’t slow down,
then scientists would have to say that the frequency of light emitted from cesium
atoms is gradually increasing for seemingly no apparent reason. That is, by sticking to
the earth-sun clock, we have trouble accounting for accelerations and retardations of
the orbital motions of the other planets compared with earth’s rotational period, and
we have trouble accounting for the simultaneous accelerations and retardations of
atomic motions such as those in cesium-133 atoms compared again with earth’s
rotational period. Our atomic theory says that these atomic processes should behave
uniformly as time goes on, so sticking with the earth-sun clock forces us accept
awkward changes in our atomic theory and in the rotations of the other planets. On the
other hand, by switching to the cesium atomic standard, these alterations are
unnecessary, the mysteries vanish, and we can readily explain the non-uniform
wobbling of the earth’s yearly revolutions by reference to the tides on the earth, the
movement of the liquid metal at the center of earth, the gravitational pull of other
planets, dust between planets, and collisions with comets. These influences affecting a
solar clock do not affect the cycles of the cesium atom.

There are two principal advantages of the cesium clock: (1) it provides a standard that
is reproducible anywhere in the universe where there is cesium, and, more importantly,
(2) the behavior of the cesium atom is relatively insulated or isolated from other
processes, especially from a comet’s bombarding the earth.

In order to keep our atomic-based calendar in synchrony with the rotations and
revolutions of the earth, say, to keep atomic-noons occurring on astronomical-noons
and ultimately to prevent Northern hemisphere winters from occurring in some future
July, we systematically add leap years and leap seconds and leap microseconds in the
counting process. These changes don’t affect the duration of a second, but they do
affect the duration of a year because, with leap years, not all years last the same
number of seconds.

Our universe has a large number of different processes that bear consistent time
relations, or frequency of occurrence relations, to each other. For example, the
frequency of a fixed-length pendulum is a constant multiple of the half life of a specific
radioactive uranium isotope; the relationship doesn’t change as time goes by (at least
not much and not for a long time). The existence of these sorts of relationships makes
our system of physical laws much simpler than it otherwise would be, and it makes us
more confident that there is something objective we are referring to with the time-
variable in those laws.

24. What does it mean for a clock to be accurate?


It’s important to distinguish accuracy from precision. If you use a bow to shoot arrows
at a target, then the shooting is precise if all the arrows cluster near a point, even if
that point is off-target. For your shooting to be accurate you need to hit the bull’s-eye.
The standard clock’s ticking is our bulls-eye. An ordinary wristwatch is considered to be
accurate if it ticks in synchrony (that is, in step) with our standard clock.
What it means for the standard clock to be accurate depends somewhat on your
philosophy of time. If you are a conventionalist, then once you pick the convention, the
standard clock can’t fail to be accurate. There may be more or less useful standards
(you would do better choosing the ticks of an atomic clock rather than the barks of your
neighbor’s dog as the standard periodic process), but usefulness isn’t a sign of truth.
The absolute theory of time, on the other hand, implies time is marching on
independent of all events, and an accurate standard clock will be in sync with this
“march.” If it is out of sync, then our standard clock won’t be telling the true time. But
since our civilization doesn’t know how to establish this synchrony, we take a very
different route to accuracy by saying the best choice for a standard clock is one that is
the most regular, and we find out which is the most regular by finding the clock that is
best at meeting the following three goals:

1. The most accurate clock will use a process that is not affected very much by
environmental conditions such as temperature, time of day, where it’s located,
the presence of dust and comets. [The standard atomic clock meets goal (a)
better than the standard astronomical clock does.]
2. Exact reproductions of the clock should stay in synchrony with each other when
environmental conditions are the same. To use the technical expression, the
reproductions should remain sufficiently congruent, i.e., more congruent than
competing clocks using a different standard.
3. The standard clock’s readings should be consistent with the Newton’s first and
second laws of motion (assuming we are in a situation where these laws should
hold so that we don’t need to deal with Einstein’s revisions of Newton’s laws). If
we run a test of those laws, and if we find that Newton’s laws are violated, then
the problem isn’t with the laws but with the clock used in our test, and we say the
clock is inaccurate, provided there are no other mistakes in the experiment. The
first person to notice requirement (c) on accuracy of clocks was Leonhard Euler
[1707-1783], a Swiss mathematician and physicist.

Author Information

Bradley Dowden
Email: dowden@csus.edu
California State University Sacramento

You might also like