Barbour, Julian B The End of Time The Next Revolution in Physics
Barbour, Julian B The End of Time The Next Revolution in Physics
Julian Barbour
Oxford New York
Athens Auckland Bangkok Bogotá Buenos Aires
Cape Town Chennai Dar es Salaam Delhi Florence Hong Kong
Istanbul
Karachi Kolkata Kuala Lumpur Madrid Melbourne Mexico City
Mumbai
Nairobi Paris São Paulo Shanghai Singapore Taipei Tokyo Toronto
Warsaw
Berlin Ibadan
9 10
Printed in the United States of America
on acid-free paper
CONTENTS
Preface
Acknowledgements
Newton’s Concepts
First Outline
Is Motion Real?
An Alternative Arena
Apparent Failure
Energy
Exploring Platonia
Historical Accidents
Funny Geometry
Introduction
Bell’s Inequalities
A Dualistic Picture
A Simple-Minded Approach
A Weil-Ordered Cosmos?
Notes
Further Reading
Bibliography
Index
BOXES
1 The Great Revolutions in Physics
3 Possible Platonias
4 Centre of Mass
12 Entangled States
ILLUSTRATION
2 Motion as an Illusion
3 Triangle Land
5 Platonia
8 Shape Space
11 Centre of Mass
15 A Spiral Galaxy
17 Potential Energy
27 Space-Time Diagram
30 Three-Spaces in Space-Time
31 Space-Time as a Tapestry
34 Actual Distribution
37 Superposition of Waves
42 Entangled States
52 Division of Platonia
Two views of the world clashed at the dawn of thought. In the great
debate between the earliest Greek philosophers, Heraclitus argued
for perpetual change, but Parmenides maintained there was neither
time nor motion. Over the ages, few thinkers have taken Parmenides
seriously, but I shall argue that Heraclitan ux, depicted nowhere
more dramatically than in Turner’s painting below, may well be
nothing but a well-founded illusion. I shall take you to a prospect of
the end of time. In fact, you see it in Turner’s painting, which is
static and has not changed since he painted it. It is an illusion of
ux. Modern physics is beginning to suggest that all the motions of
the whole universe are a similar illusion – that in this respect Nature
is an even more consummate artist than Turner. This is the story of
my book.
Snow Storm – Steamboat o a Harbour’s Mouth Making Signals in
Shallow Water, and Going by the Lead. The Author was in this Storm on
the Night the Ariel left Harwich (1842). The 67-year-old Turner
claimed that he had made the sailors bind him to the Ariel’s mast so
that he should be forced to experience the full fury of the storm.
PREFACE
You may wonder how I can preface a belief that time does not
exist with a bit of personal history. How can history be if there is no
time? That is the great question, and my answer comes at the end of
the book. Most of the book is about what evidence physics can o er
for and against the existence of time. However, in the rst part I try
to explain, in the simplest terms possible, the main issues, and to
relate them to your direct experience of time. I want to try to make
sure, if you have bought or borrowed this book, that you do not put
it down in despair, unable to understand what I am driving at. I
hope also that this introduction will encourage you to read on to the
details. Many are fascinating in their own right. Because temporal
concepts are so deeply lodged in our experience and language, I
shall often write as if time existed in the way most people think it
does. The same applies to motion. Please do not think I am being
inconsistent – I should have to use many more words to express
everything in a timeless fashion.
I have tried to make the text self-contained and accessible to any
reader fascinated by time. If you nd some parts harder then others,
please do not worry if you have to give up on them. Several non-
scientists who read a much more technical rst draft found they
could simply skim the harder parts and still pick up much of the
message. For this reason, the more technical material that is not
completely central to the story is generally put in boxes – take that
as a sign not to worry if you have di culty digesting it (though I
hope you will at least try it). Also, various digressions, of potential
interest to all readers, and genuinely technical material for
cognoscenti are to be found in the notes at the end. I suggest you
look at them after you have read each chapter. To help readers with
little or no scienti c background, the most important technical
terms appear in the Index so that you can readily locate
explanations of them in the text if necessary. Books for further
reading are also recommended.
J.B.
South Newington, March 1999
Note This printing of the book di ers from the initial hardback in
the correction of some minor errors and misprints, additions to the
bibliography and books recommended for further reading, and slight
rearrangement of the Notes to take into account new results
obtained with Niall Ó Murchadha after the book had been written.
This recent work should, if it stands up to critical examination,
strengthen my arguments that time does not exist. See especially p.
358.
ACKNOWLEDGEMENTS
I have left to the end one other important person – you, the
reader. As you will know from the Preface, I have tried throughout
my life to fund my own research and would like to continue to do
so. Every copy of this book that is bought (and borrowed from a
library) helps me in this way. Thank you, and I do hope you get
some pleasure from this book. I have enjoyed writing it. I hope to
continue popularizing the study of time and will post details on my
Website (www.julianbarbour.com) together with any signi cant
developments of which I become aware in the study of time.
PART 1
The Big Picture in Simple Terms
No doubt many people will dismiss the suggestion that time may
not exist as nonsense. I am not denying the powerful phenomenon
we call time. But is it what it seems to be? After all, the Earth seems
to be at. I believe the true phenomenon is so di erent that,
presented to you as I think it is without any mention of the word
‘time’, it would not occur to you to call it that.
One of the themes of my book is that this chasm has arisen because
physicists have deep-rooted but false ideas about the nature of space,
time and things. Preconceptions obscure the true nature of the world.
Physicists are using too many concepts. They assume that there are
many things, and that these things move in a great invisible
framework of space and time.
Space and time in their previous role as the stage of the world are
redundant. There is no container. The world does not contain things,
it is things. These things are Nows that, so to speak, hover in
nothing. Newtonian physics, Einstein’s relativity and quantum
mechanics will all be seen to do di erent things with the Nows. They
arrange them in di erent ways. What is more, the rules that govern
the universe as a whole leave imprints on what we nd around us.
These local imprints, which physicists take as the fundamental laws
of nature, reveal few hints of their origin in a deeper scheme of
things. The attempt to understand the universe as a whole by
‘stringing together’ these local imprints without a grasp of their
origin must give a false picture. It will be the at Earth writ large.
My aim is to show how the local imprints can arise from a deeper
reality, how a theory of time emerges from timelessness. The task is
not to study time, but to show how nature creates the impression of
time.
We must begin by trying to agree what time is. The problems start
already, as St Augustine found. Nearly everybody would agree that
time is experienced as something linear. It seems to move forward
relentlessly, through instants strung out continuously on a line. We
ride on an everchanging Now like passengers on a train. Each point
on the line is a new instant. But is time moving forward – and if so
through what – or are we moving forward through time? It is all very
puzzling, and philosophers have got into interminable arguments. I
shall not attempt to sort them out, since I do not think it would get
us anywhere. The trouble with time is its invisibility. We shall never
agree unless we can talk about something we can see and grasp.
Suppose that the snapshots are taken when we are witnessing lots of
things happening, say people streaming past us in a street, and that
the snapshots (either two-dimensional, as directly experienced, or
‘three-dimensional’, as explained above), once taken, are jumbled up
in a heap. A di erent person, given the heap, could relatively easily,
by examining the details in the snapshots, arrange them in the order
they were experienced. A movie can be reassembled from its
individual frames. My notion of time depends crucially on the details
that the ‘snapshots’ carry. It requires the richly structured world we
do experience.
NEWTON’S CONCEPTS
There are two remarkable things about the order in the universe:
the amount of it and the way it degrades. One of the greatest
discoveries of science, made about a hundred and fty years ago,
was the second law of thermodynamics. Studies of the e ciency
with which steam engines turn heat into mechanically useful motion
led to the concept of entropy. As originally discovered, this is a
measure of how much useful work can be got out of hot gas, say. It is
here that the arrow of time, which we know from direct experience,
enters physics. Almost all processes observed in the universe have a
directionality. In an isolated system, temperature di erences are
always equalized. This means, for example, that you cannot extract
energy from a cooler gas to make a hotter gas even hotter and chu
along in your steam engine even faster. More strictly, if you did, you
would degrade more energy than you gain and nish up worse o .
Only two ways have ever been found to explain the arrow: either
the universe was created in a highly unlikely special state, and its
initial order has been ‘degrading’ ever since, or it has existed for
ever, and at some time in the recent past it entered by chance an
exceedingly improbable state of very low entropy, from which it is
now emerging. The second possibility is entirely compatible with the
laws of physics. For example, if a collection of atoms (which obey
Newton’s laws) is con ned in a box and completely isolated, it will,
over a su ciently long period of time, visit (or rather come
arbitrarily near) all the states that it can in principle ever reach, even
those that are highly ordered and statistically very unlikely.
However, the intervals of time between returns to states of very low
entropy are stupendously long (vastly longer than the presently
assumed age of the universe), and neither explanation is attractive.
Put in its crudest form, a brain scientist who knew the state of
our brain would know our conscious state at that instant. The brain
state allows us to reconstruct the conscious state, just as musical
notes on paper can be transformed by an orchestra into music we can
hear. By the ‘state’ of a system, say a collection of atoms, scientists
usually mean the positions of all its parts and the motions of those
parts at some particular instant. It is widely assumed that conscious
states, in which, after all, we are aware of motion directly, are at the
least correlated with (correspond to) brain states that involve not
only instantaneous positions but also motions and, more generally,
change (associated with ow of electric currents or chemicals, for
example). This is a natural assumption. Our awareness of motion and
change is vivid and often exciting: think of watching gymnastics, or
the 100-metre sprint nal in the Olympic Games. We suppose that
the impression of motion must be created by some motion or change
in the brain.
Perhaps not. The brain often fools us. When we rst look at certain
drawings, they appear to represent one thing. After a while, the
image ickers and we see something di erent. The reason is well
understood: the brain processes information before we get it. We do
not see things as they are but as the brain interprets them for us.
There are very understandable reasons for this, but the fact remains
that we are often fooled by such ‘deceptions’.
TIME CAPSULES
As a rst example, we can stay within the brain but consider long-
term memory. A game we sometimes play at Christmas brings out
the importance of mutually consistent records held in structures.
Fifty events in recent world history are written down on separate
cards without dates attached. Players are divided into teams and
given the cards jumbled up. The challenge is to put them in the
correct chronological order. The only resource each team has to
attack the task is their collective long-term memories, which every
good realist (myself included) will surely agree are somehow or
other ‘hard-wired’ into their brains. How each team fares depends on
the consistency of its members’ recollections – the records in their
brains.
This example shows clearly that all we know about the past is
actually contained in present records. The past becomes more real
and palpable, the greater the consistency of the records. But what is
the past? Strictly, it is never anything more than we can infer from
present records. The word ‘record’ prejudges the issue. If we came to
suspect that the past is a conjecture, we might replace ‘records’ by
some more neutral expression like ‘structures that seem to tell a
consistent story’.
For me, two facts above all stand out from this miracle of nature.
If we discount the direct perception of motion in consciousness, all
this fantastic abundance of evidence for time and history is coded in
static con gurational form, in structures that persist. This is the rst
fact, and it is ironic. The evidence for time is literally written in
rocks. This is why I believe the secret of time is to be unravelled
through the notion of time capsules. It is also the reason why I seek
to reduce the other hard and persistent evidence for time and motion
– our direct awareness of them in consciousness – to a time-capsule
structure in our brains. If I can make such a structure responsible for
our short-term memory – the phenomenon of the specious present –
and for the actual seeing of motion, then all appearances of time will
have been reduced to a common basis: special structure in individual
Nows.
FIRST OUTLINE
I think you will agree that the Current Theory bag does match
experience quite closely. The triangles stand for each of the instants
you experience, and they follow one another continuously, just as the
instants do. By giving them to you in a bag and getting you to lay
them out in a sequence, I am giving you a ‘God’s eye’ view of history.
All its instants are, as it were, spread out in eternity as if you
surveyed them from a mountain-top. In fact, this way of thinking
about time has long been a commonplace among Christian
theologians and some philosophers, and has prompted them to claim
that time does not exist but that its instants all exist together and at
once in eternity. My claim is much stronger. I am saying that reality,
if we could see all of it, is not at all like the contents of the Current
Theory bag with its single sequence of states. It is like the contents of
the Timeless Theory bag, in which in principle all conceivable states
can be present. Nothing in it resembles our experience of history as a
unique sequence of states: that experience is usually explained by
assuming that there is a unique sequence of states. I deny that there
is such a sequence, and propose a di erent explanation for the
experience that prompts us to believe in it. The only thing the bags
have in common with our direct experience of time is the parallel
between individual triangles as models of individual instants of time.
This is where the assumption that all the structures found in the
bag come in multiple copies, and that the numbers of these copies,
which can vary very widely, are determined by a de nite timeless
rule, becomes crucial. Imagine that all the structures for which the
numbers of identical copies in the bag are large are time capsules,
while there are few copies of structures that are not time capsules.
Since the overwhelming majority of possible structures that can exist
are certainly not time capsules, any rule that does ll the bag with
time capsules will be remarkably selective, creative one might say. If,
in addition, you can nd evidence that the universe is governed by a
timeless law whose e ect is to discriminate between structures and
which actually selects time capsules with surprising accuracy, then
you might begin to take such ideas more seriously. You might begin
to see a way in which the Timeless Theory could still explain our
experience of time, and could perhaps be superior to the Current
Theory.
Now I want to give you a better feel for what a timeless universe
could be like. What we need rst is a proper way to think about
Nows.
One issue that runs through this book is this: what is the ultimate
arena of the universe? Is it formed by space and time (space-time), or
something else? This is the issue raised by Dirac’s sentence I quoted
in the Preface: ‘This result has led me to doubt how fundamental the
four-dimensional requirement in physics is.’ I believe that the
ultimate arena is not space-time. I can already begin to give you an
idea of what might come in its place.
Platonia is the arena that I think must replace space and time.
Why this should be so, how it can be done, and what physics in
Platonia is like is the meat of the book. But it is already possible to
see how di erently creation and a supposed beginning of time
appear in Platonia. Most people are ba ed that time could begin.
How many times do we hear the question, ‘But what happened
before the Big Bang?’ The question reveals the depth to which the
notion of an eternally owing time is ingrained in the psyche. This is
why I call the instants of time ‘things’, so as to break the spell, and
why I have chosen the name Platonia for our home. It is also why I
use paths as the image of history. In itself, there are no paths in
Platonia, just as there were no paths on Earth before animals made
them. The points of Platonia – the Nows – are worlds unto
themselves. No thread of time joins them up. We must think of
Newtonian-type dynamics as something that ‘paints a path’ onto the
timeless landscape of Platonia.
IS MOTION REAL?
We had a cat called Lucy, who was a phenomenal hunter. She could
catch swifts in ight, leaping two metres into the air. She was seen in
the act twice, and must have caught other victims since several times
we found just the outermost wing feathers of swifts by the back door.
Faced with facts like this, isn’t it ridiculous to claim there is no
motion?
There are two parts to my claim that time does not exist. I start
from the philosophical conviction that the only true things are
complete possible con gurations of the universe, unchanging Nows.
Unchanging things do not travel in time from Now to Now. Material
things, we included, are simply parts of Nows. This philosophical
standpoint must be matched by a physical theory that seems natural
within it. The evidence that such a physical theory exists and seems
to describe the universe forms the other part of my claim. This
section has merely made the philosophy, the notion of being, clear.
The physics, the guts of the story, is still to come.
You may think that time capsules and a brain preserved in aspic
aware of seeing motion are getting dangerously close to solipsism
and the machinations of a demon. Without anticipating the rest of
the book, an outline may still be helpful. There are only two rules of
the game: there must be an external world subject to laws and a
correspondence between it and experiences.
Apart from the fact that Newton placed the material objects of
the universe in an arena, my things are his things. They are Nows,
the relative con gurations of the universe. Newton’s Nows form a
string, brought into being by an act of creation at one end, called the
past. It is usually assumed that our experiences in some instant
re ect the structure in a short segment of the string at a point along
it. It is a segment, rather than one Now, because we see things not
only in positions but in motion. However, a single Now contains only
positional information. It seems that we need at least two Nows to
have information about changes of position.
(1) All experience we have in some instant derives from the structure
in one Now.
(3) The Nows at which the mist has a high intensity are time
capsules (they will also possess other speci c properties).
Thus, the one law of the universe that determines the mist intensity
over Platonia is timeless. The Nows and the distribution of the mist
are both static. The appearance of time arises solely because the mist
is concentrated on time capsules, and a Now that is a time capsule is
therefore much more likely to be experienced than a Now that is not.
(Please remember that this is only an outline: the detailed arguments
are still to come.)
Each Now has a mist intensity. Suppose that all the Nows
participate in a lottery, receiving numbers of tickets proportional to
their mist intensities. Nows where the mist is intense get tickets
galore, others very few. By assumption (1), conscious experience is
always in one Now. If a Now has a special structure, it is capable of
self-awareness. But is it actually self-aware? Structure in itself, no
matter how intricate and ordered, cannot explain how it can be self-
aware. Consciousness is the ultimate mystery.
The theory is still testable because only Nows with high mist
intensity (and therefore high probability) are likely to be
experienced, and such Nows have characteristic properties: above all,
they are time capsules. We can therefore test our own experiences
and see if they verify the predictions of the theory. This is something
that in principle can be settled by mathematics and observations. For
if physicists can determine or guess the structure of Platonia and
formulate the law that determines how the mist is distributed over it,
then it is simply a matter of calculation to nd out where in Platonia
the mist is most intense. If the mist is indeed concentrated on
structures that are time capsules, the theory will make a very strong
prediction – any Now that is experienced will contain structures that
seem to be records of a past of that Now. It will also contain other
characteristic structures.
Both Copernicus and Kepler believed that the universe, with the
solar system at its centre, was bounded by a huge and distant rigid
shell on which the luminous stars were xed. They did not speculate
what lay beyond – perhaps it was simply nothing. They de ned all
motions relative to the shell, which thus constituted an
unambiguous framework. Many factors, above all Galileo’s
telescopic observations in 1609 and the revival of interest in the
Greeks’ idea of atoms that move in the void, destroyed the old
cosmology. New ideas crystallized in a book that Descartes wrote in
1632. He was the rst person to put forward clearly an idea which,
half a century later, Newton would make into the most basic law of
nature: if nothing exerts a force on them, all bodies travel through
space for ever in a straight line at a uniform speed. This is the law of
inertia. Descartes never published his book because in 1633 the
Inquisition condemned Galileo for teaching that the Earth moves.
The Copernican system was central to Descartes’s ideas, and to
avoid Galileo’s fate he suppressed his book.
He did publish his ideas in 1644, in his in uential Principles of
Philosophy, but with a very curious theory of relative motion as an
insurance policy. He argued that a body can have motion only
relative to some other body, chosen as a reference. Since any other
body could play the role of reference, any one body could be
regarded as having many di erent motions. However, he did allow a
body to have one true ‘philosophical motion’, which was its motion
relative to the matter immediately adjacent to it. (Descartes believed
there was matter everywhere, so any body did always have matter
adjacent to it.) This idea let him o the Inquisition’s hook, since he
claimed that the Earth was carried around the Sun in a huge vortex,
as in a whirlpool. Since the Earth did not move relative to the
immediately adjacent matter of the vortex, he argued that it did not
move!
A century and a half passed before the issue became a hot topic
again. This raises an important issue: how could mechanics have
dubious foundations and yet ourish? That it ourished
nevertheless was due to fortunate circumstances that are very
relevant to the theme of this book. First, although the stars do move,
they are so far away that they provide an e ectively rigid
framework for de ning motions as observed from the Earth. It was
found that in this framework Newton’s laws do hold. It is hard to
overestimate the importance of this fortunate e ective xity of the
distant stars. It presented Newton with a wonderful backdrop and
convenient framework. Had the astronomers been able to observe
only the Sun, Moon and planets but not the stars (had they been
obscured by interstellar dust), Newton could never have established
his laws. Thus, scientists were able to accept Newton’s absolute
space as the true foundation of mechanics, using the stars as a
substitute for the real thing – that is, a true absolute frame of
reference. They also found that Newton’s uniformly owing time
must march in step with the Earth’s rotation, since when that was
used to measure time (in astronomical observations spanning
centuries, and even millennia) Newton’s laws were found to hold.
Once again, a substitute for the ‘real thing’ was at hand. One did not
have to worry about the foundations. Fortunate circumstances like
these are undoubtedly the reason why it is only recently that
physicists have been forced to address the issue of the true nature of
time.
The person who above all brought the issue of foundations back
to the fore was the Austrian physicist Ernst Mach, whose brilliant
studies in the nineteenth century of supersonic projectiles and their
sonic boom are the reason why the Mach numbers are named after
him. Mach was interested in many subjects, especially the nature
and methods of science. His philosophical standpoint had points in
common with Bishop Berkeley, but even more with the ideas of the
great eighteenth-century Scottish empiricist David Hume. Mach
insisted that science must deal with genuinely observable things,
and this made him deeply suspicious of the concepts of invisible
absolute space and time. In 1883 he published a famous history of
mechanics containing a trenchant and celebrated critique of these
concepts. One suggestion he made was particularly in uential.
AN ALTERNATIVE ARENA
It was six or seven years before I came to form really clear ideas.
I eventually concluded that what was needed above all was a new
arena in which to describe the universe. I arrived at the notion of
Platonia (or, as I originally called it, the relative con guration space
of the universe). The argument was quite simple. First, it is a fact
that we orient ourselves in real life by objects we actually see, not
by invisible space (see the Notes on the previous chapter). Things
are the signposts that tell us where we are. There is also the
fortunate fact that we live on the nearly rigid Earth. We can orient
ourselves by means of just a few objects xed on its surface, say
church spires when hiking in the English countryside. Always there,
the Earth provides a natural background. Motion seems to take
place in a framework. But imagine what life would be like if we
lived on a jelly sh!
The fact is that we live in a very special location. Only the tiniest
fraction of matter in the solar system, let alone the universe, is in
solid form. Imagine that we lived in an environment much more
typical of the universe – in space. To simplify things, let there be
only a nite number of objects, all in motion relative to one
another. At any instant there are certain distances between these
other objects and us. There is nothing else. In these circumstances,
what would be the natural way to answer what is always a
fundamental question: where are we? We have no other means of
saying where we are except in terms of our distances to other
objects. What is more, it would be arti cial to choose just a few of
them to locate ourselves. Why these rather than those? It would be
much more natural to specify our distances to all objects. They
de ne our position. This conclusion is very natural once we become
aware that nothing is xed. Everything moves relative to everything
else.
You may nd that this chapter requires more re ection than all
the others. You certainly do not need to grasp it all, but I hope that
you will be able to change from a way of thinking to which we have
been conditioned by the fact that we evolved on the stable surface of
the Earth to a more abstract way of thinking that would have been
forced upon us had we evolved from creatures that roamed in space
between objects moving through it in all directions. We have to learn
how to nd our bearings when the solid reassuring framework of the
Earth is not there. This is the kind of mental preparation you need to
understand the ideas Poincaré developed. In this respect, he was
smarter than Einstein.
If Mach is right, so that time is nothing but change and all that
really counts in the world is relative distances, there should be a
perfect analogue of Laplace’s scenario with a divine intelligence that
contemplates Platonia. Machian dynamics in Platonia must be about
the determination of paths in that timeless landscape. It should be
possible to specify an initial point in Platonia and a direction at that
point, and that should be su cient to determine the entire path.
Nothing less can satisfy a rational mind. The history in Figure 10
starts at the centre of Shape Space, so that there the particles form an
equilateral triangle, and set o in a certain direction. In Machian
dynamics, the initial position and initial direction (strictly in
Triangle Land, not Shape Space) should determine the complete
curve uniquely. Now we can test this idea in the real world. The
heavens provide plenty of triple-star systems, and astronomers have
been observing their behaviour for a long time. They certainly meet
the Laplace-type condition when described in Newtonian terms. But
are their motions comprehensible from a Machian point of view?
This is the question Poincaré posed.
APPARENT FAILURE
The answer is very curious. The motions are nearly but not quite
comprehensible. This can be highlighted by showing how di erent
possible Newtonian motions look when represented as curves in
Triangle Land, our model Platonia – or rather Shape Space, since this
is much easier to represent. To create a vivid picture, let us imagine
that we are holding two cardboard triangles that are slightly
di erent. These can represent the relative con gurations of three
mutually gravitating bodies at two slightly di erent instants of
Newton’s absolute time.
Figure 10 Another possible path traced by the same three particles
as in Figure 9 (I refer to them now as bodies). This history starts (or
ends if time is assumed to run the other way) at the con guration in
which the three bodies form an equilateral triangle. The shape of the
triangle changes in a de nite way at all points along the curve. If it
left the initial equilateral triangle along one of the dotted lines, that
would mean two sides of the triangle remaining equal in length
while the third changes – the equilateral triangle would become an
isosceles triangle. In fact, for the example shown, the ratios of all
three sides change. For readers used to thinking of motions in
ordinary space, this example corresponds to particles that constantly
orbit each other in a xed plane. The positions at which the curve
touches the dotted line that bound Shape Space correspond to
eclipses, when one particle is between the other two and on the line
joining them. Such a con guration is called a syzygy (that’s a nice
word to show o with). In ordinary space, the one particle passes
through the line joining the other two and comes out the other side.
But the points on the curve in Figures 9 and 10 stand for the
complete triangle, not one of the three particles. This is why the
curve approaches the syzygy frontier and then returns into the
interior of Shape Space. There are no triangles outside the syzygies!
Galileo noted that all physical e ects in the closed cabin of a ship
sailing at uniform speed on a calm sea unfold in exactly the same
way as in a ship at rest. Unless you look out of the porthole, you
cannot tell whether the ship is moving. Quite generally, in
Newtonian mechanics the uniform motion of an isolated system has
no e ect on the processes that take place within it. The left-hand
diagram in Figure 12 shows (in perspective) the triangles formed by
the three gravitating bodies in the history of Figure 10 at equal
intervals of absolute time. The individual bodies move along the
‘spaghetti’ tubes. The centre of mass moves uniformly up the z axis.
(Despite appearances, the triangles are always horizontal, i.e. parallel
to the xy plane.) The right-hand diagram has two physically
equivalent interpretations. First, it is how observers moving
uniformly to the left past the system on the left would see that
system receding behind them. Second, it is also how observers at rest
relative to the system on the left would see a system identical to that
system except for a uniform motion of the centre of mass to the right.
This is how the happenings in the cabin of Galileo’s galley would be
‘sheared’ to the right for observers standing on the shore. Depending
on the speed of the system, the centre of mass will be shifted in unit
time by di erent amounts, but the actual sequence of triangles
remains the same. This corresponds to the freedom mentioned in the
text.
Figure 12 Unlike Figures 9 and 10 (and the later Figure 14), the
lines followed by the spaghetti strands in this gure (and also Figure
13) show the tracks of the three individual particles in space. This is
why there are three strands and not a single curve. It will help you a
lot if you can get used to thinking about these two di erent ways of
representing one and the same state of a airs. Here we see
individual particles moving in absolute space. In Figures 9, 10 and
14 we ‘see’ (in our mind’s eye) the ‘world’ or ‘universe’ formed by
the three particles moving in Platonia.
There are only four freedoms that remain. Having placed the
centre of mass of the second triangle at some position, we can
change its orientation (three freedoms). We can also change the
amount of Newton’s absolute time that elapses between the instants
at which the three bodies occupy the two positions (one freedom, the
fourth). If the time di erence is shortened, this means that the bodies
travel farther in less of Newton’s time – that is, they are moving
faster initially. In fact, since the motion of the centre of mass does
not matter, we can keep it xed and change only the orientation.
Now, at last, we come to something that does matter. Both these
changes – in the time di erence and in the relative orientation –
have dramatic consequences, which are illustrated in Figures 13 and
14.
But these are our arbitrary choices. Once we have chosen two
triangles, nothing about the triangles in themselves gives any hint as
to how we should make the choices. Leibniz formulated two great
principles of philosophy that most scientists would adhere to. The
rst is the identity of indiscernibles: if two things are identical in all
their attributes, then they are actually one. They are the same thing.
The second, which we have already met, is the principle of su cient
reason: every e ect must have a cause. There must be some real
observable di erence that explains di erent outcomes.
Before we look at the one possibility that can resolve this puzzle,
it is worth considering how the four freedoms that do count show up
in practice. We shall then be able to see what a great discovery
Newton’s invisible framework was. We start with the twists.
When I was a boy, there was only one sport at which I was any good:
the high jump. One year I went on a training course at the athletics
ground in Oxford. We were introduced to angular momentum and
how it could be exploited to improve the jump. As tallest of the
young hopefuls, I was chosen to give a demonstration. The instructor
made me lie on my side, arms and legs outstretched, on a small
bench turntable. He started to rotate it slowly and asked, ‘If you pull
yourself into a crouched position, what will happen?’ I knew, I was
studying physics: ‘Angular momentum will be conserved, and the
turntable will spin faster.’ ‘Right,’ he said, ‘do it.’ Proudly, I pulled in
my arms and legs with vigour. The e ect was frightening. The
turntable whizzed around so fast that I panicked, tried to get o , and
was thrown onto the oor. I escaped with bruises. I am still kicking
myself, not about the accident, but because I did not stay on another
day. I would have seen Roger Bannister run the rst four-minute
mile.
Angular momentum is a kind of net spin about a xed axis. To
calculate it for the Earth, you multiply the mass of each piece of
matter in the Earth by its perpendicular distance from the rotation
axis and the speed of its circular motion about the axis. The Earth’s
total angular momentum is the sum of the contributions of all the
pieces. Clockwise and anticlockwise motions count oppositely. A jet
plane ying round the world in the opposite sense to the daily
rotation contributes with the opposite sign.
The fact that all the planets move in the same direction around
the Sun in nearly coincident planes is thus a remote consequence of
the relatively modest initial net spin of the primordial dust cloud. We
see the result in the sky, since all the celestial wanderers – the Sun,
Moon and planets – follow much the same track against the
background of the stars. Ironically, Newton underestimated the
power of his own laws. He could not bring himself to believe that the
solar system had arisen naturally. ‘Mere mechanical causes’, he said,
‘could not give birth to so many regular motions.’ He asserted that
‘this most beautiful system’ could only have proceeded ‘from the
counsel and dominion of an intelligent and powerful Being’. One
wonders what Newton would have made of the modern pictures of
Saturn and its rings (Figure 16). Of all the images created in the
heavens by gravity and the invariable plane, this is surely the most
perfect.
For three centuries, the best explanation for phenomena like the
rings of Saturn has remained Newton’s: inertia, the inherent
tendency of all objects to follow straight lines in the room-like arena
of absolute space. If these are accepted, then the rings of Saturn,
tops, frisbees and all the other manifestations of angular momentum
can be explained. However, Newton’s account is not so much an
explanation as a statement of facts in need of explanation. Since it is
always matter that we actually see, should we not try to account for
these things without the mysterious intermediaries of absolute space
and time? Before we attack this problem, we need to consider energy
and, in the next chapter, clocks and the measurement of time.
Figure 16 Saturn and its rings.
ENERGY
WHERE IS TIME?
at t = 1, distance fallen = 1,
at t = 2, distance fallen = 1 + 3 = 4,
at t = 3, distance fallen = 1 + 3 + 5 = 9,
at t = 4, distance fallen = 1 + 3 + 5 + 7 = 16,…
Nearly two thousand years ago, astronomers knew that some motions
are better than others as measures of time. This they discovered
experimentally. For the early astronomers, there were two obvious
and, on the face of it, equally good candidates for telling time. Both
were up in the sky and both had impressive credentials. The stars
made the rst clock, the Sun the second.
The stars remain xed relative to each other and de ne sidereal
time. Any star can be chosen as the ‘hand’ of the stellar clock: one
merely has to note when it is due south. The stellar clock then ticks
whenever that star is due south (i.e. when it crosses the meridian).
Fractions of the ‘tick unit’ are measured by its distance from the
meridian. A mere glance at the night sky could tell the ancient
astronomers the time to within a quarter of an hour. With some care,
times could be told to a minute. There is something wonderful about
this great clock in the sky. It was a unique gift to the astronomers.
The discoveries that culminated with Kepler’s laws of planetary
motion, and many more made until well into the twentieth century,
are unthinkable without it. No other phenomenon in nature could
match it for convenience and accuracy. In millennia it has lost a few
hours.
But there is a rival – the Sun. It de nes solar time. This is the
clock by which humanity and all other animals have always lived.
The principle is the same: it is noon when the sun crosses the
meridian. You don’t even have to be an astronomer to tell the time
by this clock; a sundial will do.
The rst di erence between sidereal and solar time arises from one
of the three laws discovered by Kepler that describe the motion of
the planets. The Sun’s apparent motion round the ecliptic is, of
course, the re ection of the Earth’s motion. But, as Kepler
demonstrated with his second law, that motion is not uniform. For
this reason, the Sun’s daily eastward motion varies slightly during
the year from its average. The di erences build up to about ten
minutes at some times of the year.
Since the Sun is much more important for most human a airs
than the stars, how did the astronomers persuade governments to
rule by the stars? What makes the one clock better than the other?
The rst answer came from the Moon and eclipses. Astronomers have
always used eclipse prediction to impress governments. By around
140 BC, Hipparchus, the rst great Greek astronomer, had already
devised a very respectable theory of the Moon’s motion, and could
predict eclipses quite well.
They also explain the meaning of duration and the statement that
a second today is the same as a second tomorrow. Duration is
reduced to distance. If today or tomorrow any one of the ‘hands’ of
the inertial clock moves through the same distance, then we can say
that the ‘same amount of time’ has passed. The extra time dimension
is redundant: everything we need to know about time can be read o
from distances. But note how special is the distance that leads to a
meaningful de nition of duration. Any change of distance ‘labels’ the
instants of time. In statements like ‘Particle A hits B when C is ve
metres from D’, ‘ ve metres’ identi es an instant of time – it labels
the instant. That is su cient for history. However, the obvious
changes of distance – those between particles – do not lead to a
sensible de nition of duration. The secrets of time are rather well
hidden.
The same thing can be done for any number of bodies. Their
relative con gurations will correspond to di erent points along a
curve in the corresponding Platonia. To lay out ‘marks of equal
intervals of time’ on it, we have to go through the same procedure
with the computer, telling it to nd a framework and a time in which
the bodies do satisfy Newton’s laws. Only two facts about this
process are signi cant. First, because all the bodies interact, all their
positions must be used if the ‘time marks’ are to be found. To tell the
time by such a clock, we need to know where all its bodies are. Time
cannot be deduced from a small number of them, unlike inertial
time; the clock has as many hands as the system has bodies. Second,
no matter how many bodies there are in a system, the data in just
two snapshots are never enough to nd the spaghetti sculpture in
absolute space and construct a clock. We always need at least some
data from a third snapshot. As we have seen, this ‘two-and-a-bit
puzzle’ is the main – indeed the only – evidence that absolute space
and not Platonia is the arena of the universe.
You might think that this is all far removed from practical
considerations. It is true that scientists have learned to make
extremely accurate clocks using atomic phenomena. But this is a
comparatively recent development. Before then, astronomers faced a
tricky situation, which is worth recounting.
In the light of this, let us think again about Galileo’s ball rolling
across the table in Padua. Snapshots of the ball alone were not
su cient to tell what would happen when it rolled over the edge. It
seemed inconceivable that the ball’s path could be determined by the
little bit of water owing from a tank used to tell the time. For such
reasons as this, Newton rejected speed relative to any one motion as
a fundamental concept and invoked instead speed relative to an
abstract time. However, if we conceive the universe as a single
dynamical entity, the abstract time becomes redundant. The speed of
Galileo’s ball that determines which parabola it will trace is its speed
as measured by the totality of motions in the universe. This explains
why some motions are distinguished from others for timekeeping.
They are those that march in step with the cosmic clock, the unique
true measure of time. This time is the distillation of all change. High
noon is a con guration of the universe.
Soon there developed the idea that the laws of motion – and thus
the behaviour of the entire universe – could be explained by an
optimization principle. Leibniz, in particular, was impressed by
Fermat’s principle and was always looking for a reason why one
thing should happen rather than another. This was an application of
his principle of su cient reason: there must be a cause for every
e ect. Leibniz famously asked why, among all possible worlds, just
one should be realized. He suggested, rather loosely, that God – the
supremely rational being – could have no alternative but to create
the best among all possible worlds. For this he was satirized as Dr
Pangloss in Voltaire’s Candide. In fact, in his main philosophical
work, the Monadology, Leibniz makes the more defensible claim that
the actual world is distinguished from other possible worlds by
possessing ‘as much variety as possible, but with the greatest order
possible’. This, he says, would be the way to obtain ‘as much
perfection as possible’.
The rst idea Bruno and I developed had several interesting and
promising properties. Above all, it showed that a mechanics of the
complete universe containing only relative quantities and no extra
Newtonian framework could be constructed. Hitherto, most people
had thought this to be impossible. Just as Mach had suspected, the
phenomenon that Newton called inertial motion in absolute space
could be shown to arise from motion relative to all the masses in the
universe. We also showed that an external time is redundant.
However, besides the desirable features we obtained e ects which
showed that the theory could not be right. While the universe as a
whole could create the experimentally observed inertial e ects that
we wanted, the Galaxy would create additional e ects, not observed
by astronomers, that ruled out our approach.
I mention these things because the next idea that Bruno and I
tried seems to me just as natural as our rst idea, if one approaches
the problems of describing motion and change with an open mind. It
does, however, seem very di erent from the present paradigm,
which has become deep-rooted with the long hegemony of
Newtonian ideas, which were only partly changed by Einstein.
Although, as we shall see, our second idea is actually built into
Einstein’s theory at its very heart, within the context of classical
physics it merely provides a di erent perspective on that theory.
However, for the study of quantum e ects it does represent a
genuine alternative, and the attempt to create a quantum theory of
the universe may force its adoption, alien though it may appear to
many working scientists.
EXPLORING PLATONIA
Let me now explain this second idea. So far, I have explained only
what the points of Platonia are. Each is a possible relative
arrangement, a con guration, of all the matter in the universe. If
there are only three bodies in it, Platonia is Triangle Land, each
point of which is a triangle (Figures 3 and 4). Can we somehow say
‘how far apart’ any two similar but distinct triangles are? If so, this
will de ne a ‘distance’ between neighbouring points in Triangle
Land, and just as mathematicians seek geodesics using ordinary
distances on curved surfaces, we can start to look for geodesics in
Platonia. If we can nd them, they will be natural candidates for
actual histories of the universe, which we have identi ed as paths in
Platonia. They will be Machian histories if the ‘distance’ between
any two neighbouring points in Platonia is determined by their
structures and nothing else, and we shall not need to suppose that
they are embedded in some extra structure like absolute space.
In Figure 21, triangle ABC is one point in Triangle Land, and the
slightly di erent triangle A*B*C* is a neighbouring point. A
‘distance’ between them can be found in many ways, but one of the
simplest is the following. Imagine that ABC is held xed, and
A*B*C* is placed in any position relative to it. This creates
‘distances’ AA*, BB*, CC* between the corresponding vertices, at
which we suppose there are bodies of masses a, b, c. We form a ‘trial
distance’ d by taking each mass and multiplying it by the square of
the corresponding distance, adding the results and taking the square
root of the sum. Thus
You can see directly how absolute space and time are created out
of timelessness. Take some point on one of the Machian geodesics in
Platonia; it is a con guration of masses. Take another point a little
way along the geodesic; it is a slightly di erent con guration.
Without any use of absolute space and time, using just the two
con gurations, you can bring the second into the position of best
matching relative to the rst. You can then take a third
con guration, a bit farther along the path, and bring it into its best
matching position relative to the second con guration. You can go
along the whole path in this way. The entire string of con gurations
is oriented in a de nite position relative to the rst con guration.
What looks like a framework is created, but it is not a pre-existing
framework into which the con gurations of the universe are slotted:
it is brought into being by matching the con gurations.
Nevertheless, we get something like the Newtonian picture in Figure
1, except that we do not as yet have the ‘spacings in time’.
But this too emerges from the Machian theory. In the equations
that describe how the objects move in the framework built up by
best matching, it is very convenient to measure how far each body
moves by making a comparison with a certain average of all the
bodies in the universe. The choice of the average is obvious, and
simpli es the equations dramatically. No other choice does the
trick. For this reason it needs a special name; I shall call it the
Machian distinguished simpli er. It is directly related to the quantities
used to determine the geodesic paths in Platonia. To nd how much
it changes as the universe passes from one con guration to another
slightly di erent one, it is necessary only to divide their intrinsic
di erence by the square root of minus the potential. (The action, by
contrast, is found by multiplying it by the same quantity.) When this
distinguished simpli er is used as ‘time’, it turns out that each
object in the universe moves in the Machian framework described
above exactly as Newton’s laws prescribe. Newton’s laws and his
framework both arise from a single law of the universe that does not
presuppose them.
HISTORICAL ACCIDENTS
The one exception with which Lorentz had to contend was the
famous Michelson-Morley experiment performed with great accuracy
in 1887. Based on interference between light beams moving in the
direction of the Earth’s motion and at right angles to it, it was
designed to measure the change in the speed of the Earth’s motion
through the aether over the course of a year. Its accuracy was
su cient to detect even only one-hundredth of the expected e ect.
But nothing was observed. It was a great surprise, and very puzzling.
But such signals were unknown in Einstein’s time, and his theory
would show that they could not exist. He therefore used the best
substitute – light. This completely changed things. Light was to be
analysed in a framework that light itself created, so the problem
became self-referential. It might seem that Einstein cheated, making
up the rules as he went along to ensure that he won the game.
However, he was simply confronting a fact of life: laws of nature will
be meaningful only if they relate things that can actually be
observed. We live inside, not outside, the universe, and to
synchronize distant clocks we have no alternative to the physical
means available to us. Einstein’s hunch that we should use light
because it would turn out to be the fastest medium available in
nature has so far been totally vindicated.
The magical touch was that his choice was the most natural thing
to do – in the theory of an aether and in the context of the relativity
principle. Given their apparent irreconcilability, his subsequent
demonstration of their compatibility was a coup. It also showed that
there was something inevitable about the result.
For all that, general relativity does contain, hidden away in its
mathematics (as I have already indicated), a theory of duration and
the spatiotemporal framework. However, this did not come to light
for many decades and even now is not properly appreciated. How
this came about, and an account of the ‘hidden dynamical core’ of
general relativity, are the subject of the next chapters.
The views of space and time which I wish to lay before you
have sprung from the soil of experimental physics, and
therein lies their strength. They are radical. 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.
Einstein, Minkowski and others were able to show that all the
laws of nature known in their time (except initially for gravitation)
either already had a form that was exactly the same in all Lorentz
frames or could be relatively easily modi ed so that they did. Even
though the modi cations were relatively easy once the idea was
clear, their implications, including Einstein’s famous equation E =
mc2 (a prediction at that time), were mostly very startling.
Minkowski, like Einstein and Poincaré, made a strong prediction
that all laws of nature found in future would accord with the
relativity principle, and emphasized that the guiding principle for
nding such laws was to treat time exactly as if it were space.
The block universe picture is in fact close to my own, but the idea
that Nows have no role at all to play in physics, and must be
replaced by point-like events, would destroy my programme.
However, it is only absolute simultaneity that Einstein denied.
Relative simultaneity was not overthrown.
FUNNY GEOMETRY
Each event has a light cone, but only O’s is shown. Relativity
di ers from Newtonian theory mainly through the light cone and its
associated distinguished speed c, which is a limiting speed for all
processes. Light plays a distinguished role in relativity simply
because it has that speed. No material object can travel at or faster
than it. If a material object passes through O, its world line must lie
somewhere inside the light cone, for example OA in Figure 27.
There are several important things about this result. Einstein had
shown that observers moving relative to each other would not agree
about distances and times between pairs of events. However,
Minkowski found something on which they will always agree.
Measurements of the space-like separation (by a rod) and the time-
like separation (by a clock) of the same two events O and A can be
made by observers moving at any speed. They will all disagree about
the results of the separate measurements, but they will all nd the
same value for the square of the time-like separation minus the
square of the space-like separation. It will always be equal to the
square of the time-like separation, called the proper time, of the
unique observer for whom O and A are at the same space position.
This result created a sensation. Space and time, like rods and clocks,
seem to have completely di erent natures, but Einstein and
Minkowski showed that they are inseparably linked.
What is more, Minkowski showed that it is very natural to regard
space and time together as a kind of four-dimensional country in
which any two points (events in space-time) are separated by a
‘distance’. This ‘distance’, found by measurements with both rods and
clocks, is to be regarded as perfectly real because everyone will agree
on its value. In fact, Minkowski argued that it is more real than
ordinary distances or times, since di erent observers disagree on
them. Only the ‘distance’ in space-time is always found to be the
same. But it is a novel distance – positive for the time-like OA in
Figure 27, zero for the light-like OF and negative for the space-like
OC. (It is a convention, often reversed, to make time-like separations
positive and space-like ones negative. What counts is that they have
opposite signs. Also, if the units of space and time are not chosen to
make the speed of light c equal to 1, the square of the space-time
‘distance’ becomes (cT)2 – X2.)
His 1905 paper killed the idea that uniform motion relative to
any kind of absolute space or aether could be detected. But Newton
had based his case for absolute space on the detection not of uniform
motion, but of acceleration. In 1933, Einstein admitted that in 1905
he had wanted to extend the relativity principle to accelerated as
well as uniform motion, but could not see how to. The great
inspiration – ‘the happiest thought of my life’ – came in 1907 when
he started to consider how Newtonian gravity might be adapted to
the framework of special relativity. He suddenly realized the
potential signi cance of the fact, noted by Galileo and con rmed
with impressive accuracy by Newton, that all bodies fall with exactly
the same acceleration in a gravitational eld.
By 1907 Einstein was also able to show that gravity must de ect
light. Both his early predictions, made precise in his fully developed
theory, have been con rmed with most impressive accuracy in recent
decades. However, Einstein saw his early predictions merely as
stepping stones to something far grander. The equivalence principle
persuaded him that inertia (i.e. the tendency of bodies to persist in a
state of rest or uniform motion) and gravity, which Newton and all
other physicists had regarded as distinct, must actually be identical
in nature. He started to look for a conceptual framework in which to
locate this conviction. At the same time, he saw a great opportunity
to abolish not only the aether but also all vestiges of absolute space.
So far he had managed to achieve two steps in this process by
showing that uniform motion and uniform acceleration could not
correspond to anything physically real in the world. However, much
more general motions could be imagined. Einstein aimed to show
that the laws of nature could be expressed in identical form whatever
the motion of the frame of reference.
The only justi cation for the distinguished systems that appeared
in Newtonian dynamics and special relativity was the law of inertia.
But the equivalence principle had opened up the possibility of
unifying inertia and gravity. This insight sustained Einstein in his
long search for general relativity. His contemporaries would all have
been content simply to nd a new law of gravity. He was after
something sublime.
Many more things could be said about general relativity and its
discovery. However, what I want to do now is identify the aspects of
the theory and the manner of its discovery that have the most
bearing on time.
For all that, space-time does have a special, sinewy structure that
needs to be taken into account. Distinguished coordinate systems still
feature in the theory. This is because the theory of measurement and
the connection between theory and experiment is very largely taken
over from special relativity. In fact, much of the content of general
relativity is contained in the meaning of the ‘distance’ that exists in
space-time. This is where the analogy between space-time and a
landscape is misleading. We can imagine wandering around in a
landscape with a ruler in our pocket. Whenever we want to measure
some distance, we just sh out the ruler and apply it to the chosen
interval. But measurement in special relativity is a much more subtle
and sophisticated business than that. In general, we need both a rod
and a clock to measure an interval in space-time. Both must be
moving inertially in one of the frames of reference distinguished by
that theory, otherwise the measurements mean nothing. The theory
of measurement in general relativity simply repeats in small regions
of space-time what is done in the whole of Minkowski space-time in
special relativity. No measurements can be contemplated in general
relativity until the special structure of distinguished frames that is
the basis of special relativity has been identi ed in the small region
in which the measurements are to be made.
a boy who nds for the rst time a ripe horse-chestnut with
the outer shell intact. Cherishing the golden and curiously
shaped object, he might take it home, quite unaware of the
shiny brown and perfectly smooth conker ready to spring
from the shell on application of a little directed pressure.
That was Kepler’s fate: he died without an inkling of what his
nut really contained.
But it never entered his head to ask how the device actually
worked. He died only half aware of the miracle he had created.
CHAPTER 11
General Relativity: The Timeless Picture
Strange as it may seem, general relativity was little studied for about
forty years. This was not for want of admiration, for it was soon
recognized as a supreme achievement. Con rmation of the predicted
bending of starlight near the Sun by Arthur Eddington’s eclipse
expedition in 1919, communicated by telegram to The Times, made
Einstein into a world celebrity overnight. The problem was that there
seemed to be little one could do except wonder at the miracle of the
theory he had created.
This is what Dirac and ADM set out to establish. The answer was
manifestly a surprise for Dirac at least, since it led him to make the
remarkable statement quoted in the Preface. They found that if
general relativity is to be cast into a dynamical form, then the ‘thing
that changes’ is not, as people had instinctively assumed, the four-
dimensional distances within space-time, but the distances within
three-dimensional spaces nested in space-time. The dynamics of
general relativity is about three-dimensional things: Riemannian
spaces.
To connect this with the topics of Part 2, let me tell you about the
work that Bruno Bertotti and I did after the work described there. We
began to wonder whether we could be more ambitious and construct
not merely a non-relativistic, Machian mechanics, but perhaps an
alternative to general relativity. At the time, we believed that
Einstein’s theory did not accord with genuine Machian principles.
Experimental support for it was beginning to seem rather convincing,
but tiny e ects have often led to the replacement of a seemingly
perfect theory by another with a very di erent structure. We were
aware of quite a lot of the work of Wheeler and ADM, and various
arguments persuaded us that the geometry of three-dimensional
space might well be Riemannian, possess curvature and evolve in
accordance with Machian principles. We wanted to nd a Machian
geometrodynamics, which we did not think would be general
relativity. The rst task was to select the basic elements of such a
theory. What structures should represent instants of time and be the
points of the theory’s Platonia?
The issue came into clear focus for me in 1980. In April of that
year, Karel gave a memorable review talk at an international
conference in Oxford, during which I had an opportunity to discuss
with him the ideas that Bruno and I were developing. He invited me
to come to Salt Lake City, which I did in the late fall, just in time to
see the pale gold of the aspens in the Wasatch mountains. Getting to
know Utah and the magni cent deserts of the western United States
has been a great bonus from the study of physics for me and my
family. But as this is a book about physics, not travel, I had better
not digress.
On the other hand, a truly inspired genius might just have hit on
one further condition. Let dynamics do all those things with
whatever three-dimensional entities it may care to start from. But let
there be one supreme overarching principle, an even deeper unity.
All the three-dimensional things are to be, simultaneously with all
their dynamical properties, mere aspects of a higher four-
dimensional unity and symmetry.
By the end of the nineteenth century, the evidence for the wave
theory of light was very strong. However, it was precisely the failure
of light, as electromagnetic radiation, to behave in all respects in a
continuous wavelike manner that led rst Max Planck in 1900 and
then Einstein in 1905 to the revolutionary proposals that eventually
spawned quantum mechanics. A problem had arisen in the theory of
ovens, in which radiation is in thermal equilibrium with the oven
walls at some temperature. Boltzmann’s statistical methods, which
had worked so well for gases, suggested that this could not happen,
and that to heat an oven an in nite amount of energy would be
needed. The point is that radiation can have any wavelength, so
radiation with in nitely many di erent wavelengths should be
present in the oven. At the same time, the statistical arguments
suggested that, on average, the same nite amount of energy should
be associated with the radiation when in equilibrium. Therefore
there would be an in nite amount of energy in the oven – clearly an
impossibility. Baking ovens broke the laws of physics! Planck was
driven to assume that energy is transferred between the oven walls
and the radiation not continuously but in ‘lumps’, or ‘quanta’.
Most people are familiar with the speed of light, which goes
seven times round the world in a second or to the Moon and back in
two and a half seconds. The smallness of Planck’s constant is less
well known. Comparison with the number of atoms in a pea brings
it home. Angular momentum is an action and can be increased only
in ‘jerks’ that are multiples of h. Suppose we thread a pea on a string
30cm long and swing it in a circle once a second. Then the pea’s
action is about 1032 times h. As we saw, the atoms in a pea,
represented as dots a millimetre apart, would comfortably cover the
British Isles to a depth of a kilometre. The number 1032, represented
in the same way, would ll the Earth – not once but a hundred
times. Double the speed of rotation, and you will have put the same
number of action quanta into the pea’s angular momentum. It is
hardly surprising that you do not notice the individual ‘jerks’ of the
hs as they are added.
The idea of light quanta was very daring, since a great many
phenomena, above all the di raction, refraction, re ection and
dispersion of light, had all been perfectly explained during the
nineteenth century in terms of the wave hypothesis and associated
interference e ects. However, Einstein pointed out that the intensity
distributions measured in optical experiments were invariably
averages accumulated over nite times and could therefore be the
outcome of innumerable ‘hits’ of individual light quanta. Then
Maxwell’s theory would correctly describe only the averaged
distributions, not the behaviour of the individual quanta. Einstein
showed that other phenomena not belonging to the classical
successes of the wave theory could be explained better by the
quantum idea. He explained and predicted e ects in ovens, the
generation of cathode rays by ultraviolet radiation (the
photoelectric e ect), and photoluminescence, all of which de ed
classical explanation. It was for his quantum paper, not relativity,
that Einstein was awarded the 1921 Nobel Prize for Physics.
The great mystery was how light could consist of particles yet
exhibit wave behaviour. It was clear to Einstein that there must be
some statistical connection between the positions of the conjectured
light quanta and the continuous intensities of Maxwell’s theory.
Perhaps it could arise through signi cantly more complicated
classical wave equations that described particles as stable,
concentrated ‘knots’ of eld intensity. Maxwell’s equations would
then be only approximate manifestations of this deeper theory.
Throughout his life, Einstein hankered after an explanation of
quantum e ects through classical elds de ned in a space-time
framework. In this respect he was surprisingly conservative, and he
famously rejected the much simpler statistical interpretation
provided for his discoveries by the creation of quantum mechanics
in the 1920s.
During the next decade the Bohr model was applied to more and
more atoms, often but not always with success. It was clearly ad
hoc. The need for an entirely new theory of atomic and optical
phenomena based on consistent quantum principles became ever
more transparent, and was keenly felt. Finally, in 1925/6 a
complete quantum mechanics was formulated – by Werner
Heisenberg in 1925 and Erwin Schrödinger in 1926 (and called,
respectively, matrix mechanics and wave mechanics). At rst, it
seemed that they had discovered two entirely di erent schemes that
miraculously gave the same results, but quite soon Schrödinger
established their equivalence.
Heisenberg’s scheme, or picture, is based on abstract algebra and
is often regarded as giving a truer picture. In the form in which
quantum theory currently exists, it is more exible and general.
Unfortunately, it is rooted in abstract algebra, making it very
di cult to describe in intuitive terms. I shall therefore use the
Schrödinger picture. Luckily, this will not detract from what I want
to say. In fact, one of the main ideas I want to develop is that the
Schrödinger picture is actually more fundamental than the
Heisenberg picture, and is the only one that can be used to describe
the universe quantum-mechanically. Many physicists will be
sceptical about this, but perhaps this is because they study
phenomena in an environment and do not consider how local
physics might arise from the behaviour of the universe as a whole.
Although it was now clear that both light and electrons exhibited
wave-particle duality, there were important di erences between
them. A brief description of the picture as it now appears will help.
All particles are associated with elds, and can be described as
excitations of those elds. To get some idea of what this means, we
can liken the particles to water waves, which are excitations of
undisturbed water. However, the analogy is only partial. The classic
example of a particle associated with a wave is the photon, which is
an excitation of the Maxwell eld. Fields and associated particles of
di erent kinds exist. There are elds described by a single number
at each point, called scalar elds, and vector elds, which are
described by three numbers. Scalar elds represent a simple
intensity, while the vector elds – such as Maxwell’s eld – are a
kind of ‘directed’ intensity. In general relativity we also encountered
tensors. Mathematically, scalar, vector and tensor elds belong to
one family and obey the same kind of rule under rotations of the
coordinate system. In particular, after one rotation they return to
the values they had before. However, in 1927 yet another
sensational quantum discovery was made, this one by Dirac. He
found a quite di erent family of elds, called spinor elds, which are
associated with electrons and protons (as well as many other
particles). In their case, one rotation of the coordinate system brings
them back to minus the value they had before, and two rotations are
needed to restore their original value. Dirac found spinors by trying
to make the newly discovered quantum principles compatible with
relativity, and achieved a spectacular success even though it was
subsequently found that his arguments were not totally compelling.
However, the main point is that electrons are associated with a
spinor eld, photons with a vector eld.
I am not going to make any attempt to discuss this work, nor will
I try to explain the connection between a particle and its associated
eld. If a theory of everything is found, it may well change the
framework of physics. We may nd ourselves in a quite new arena
and have to change our ideas about space and time yet again.
However, as of now I believe we can glimpse the outlines of an
arena large enough to accommodate not only the present ‘zoo’ but
also whatever entities some putative theory of everything will come
up with. The arena I have in mind is vast and timeless. I see it not as
a rival to the theory of everything, but as a general framework in
which such a theory can be formulated.
INTRODUCTION
Actually, there are two mists because the wave function, being
complex, contains two numbers, which are its two components. I shall
call them the red mist and green mist, respectively. I shall also
introduce a third number, calling it the blue mist. The intensity of this
third mist is determined by the two primary components as the sum
of the squares of the red and green intensities. This is the mist
mentioned in the early chapters. Those in the know will recognize
the three mists as the real and imaginary parts of the wave function
and the square of its amplitude.
The prominence that I give to these mists could be regarded by
most theoretical physicists (above all Dirac and Heisenberg, were
they still alive) as a one-sided, if not to say distorted and naive
picture of quantum mechanics. The mists (as opposed to things called
operators) are not particularly appropriate for talking about most
quantum experiments currently performed in laboratories. However,
the experiment I have in mind is not done in a laboratory. It is what
the universe does to the instants of time. For this experiment, the one
that really counts, I think the language of mists is appropriate. Those
who disagree might have second thoughts if they really started to
think of how inertial frames and duration arise. I come back to these
issues later.
The next mystery is the collapse of the wave function. Just before
the particle hits the screen, its ψ can be spread out over a large
region. What happens to ψ when the particle is suddenly found
somewhere? The standard answer is that the wave is instantaneously
annihilated everywhere except where the particle is now known to
be.
The striking thing about this situation is that the probability for
the position of the particle, given by the sum of the squares of the
red and green intensities, is completely uniform in space. The reason
is that for two sinusoidal waves displaced by a quarter of a
wavelength, this sum is always 1 if the wave’s amplitude (its height
at the peaks) is 1. This is a consequence of the well-known
trigonometric relation sin2A + cos2A = 1, which itself is just
another expression of Pythagoras’ theorem. Thus, for a particle in
this state, we have absolutely no information about its position, but
we do know that it has a de nite momentum.
Figure 40 Like Figure 39, this shows nine di erent con gurations of
two particles (black and white triangles) on a line and the points
corresponding to them in the con guration space Q (on which a grid
has been drawn). A possible distribution of the intensity of the blue
probability mist is shown as the height of a surface over Q in the top
part of the gure (you are seeing the surface in perspective from
above, and rotated). In the state of the system shown here, the
probabilities for con gurations 4, 6 and 9 are high, while 5 has a
very low probability.
Figure 41 (a) The e ect of measuring, for the probability density of
Figure 40, the position of the particle represented by the horizontal
axis, and nding that it lies in the interval on which the vertical strip
stands. All the wave function outside the strip is instantaneously
collapsed.
This is only the start. We can select from a menu of di erent
kinds of measurement. For example, we can opt to nd the position
of only one particle, which has remarkable implications for what we
can say about the other one. Suppose rst that we measure the
position of just one of the particles. According to the quantum rules,
this instantaneously collapses the wave function from its original
two-dimensional ‘cloud’ to a one-dimensional pro le (Figure 41).
The point is that we now know the position of one particle to within
some small error, so none of the wave function outside the narrow
strip is relevant any longer. It is annihilated. If the particle whose
position is measured is represented along the horizontal axis, only a
vertical strip of ψ survives (Figure 41(a)); if the position of the other
particle is measured, only a horizontal strip survives (Figure 41(b)).
Figure 41 (b) The same for a position measurement of the other
particle.
Much more interesting is the entangled state at the top, for which
the horizontal and vertical pro les are not identical. Figure 42 shows
two pro les that result from exact position measurement of particle
1. They give very di erent probability distributions for particle 2: the
gain in information about particle 2 is considerable. This is typical of
quantum mechanics, since virtually all wave functions are entangled
to a greater or lesser extent.
I feel sure that Bohr got closer to the truth than Einstein.
However, Bohr too adopted a stance that I believe is ultimately
untenable. He insisted that it was wrong to attempt to describe the
instruments used in quantum experiments within the framework of
quantum theory. The classical world of instruments, space and time
must be presupposed if we are ever to talk about quantum
experiments and communicate meaningfully with one another. Just
as Schrödinger made his Kantian appeal to space and time as
necessary forms of thought, Bohr made an equally Kantian appeal to
macroscopic objects that behave classically. Without them, he
argued, scienti c discourse would be impossible. He is right in that,
but in the nal chapters I shall argue that it may be possible to
achieve a quantum understanding of macroscopic instruments and
their interaction with microscopic systems. Here it will help to
consider why Einstein thought the way he did.
The lent properties are the building blocks of both classical and
quantum mechanics. Classically, each particle has a unique set of
them, de ning the state of each particle at any instant. This is the
ideal to which realists like Einstein aspire. The lent properties also
occur in quantum mechanics. They are generally not the state itself,
but superpositions of them are. If a quantum system is considered in
isolation from the instruments used to study it, its basic elements still
derive from a Newtonian ontology. This is what misled EPR into
thinking they could outwit Bohr. Einstein’s defeat by Bohr is a clear
hint that we shall only understand quantum mechanics when we
comprehend Mach’s ‘overpowering unity of the All’.
BELL’S INEQUALITIES
Right at the start, Everett stated that ‘The wave function is taken
as the basic physical entity with no a priori interpretation.’ He aimed
to show that the interpretation of the theory emerges from ‘an
investigation of the logical structure of the theory’. This aim, coupled
with his insistence that the wave function is the only thing that
exists, creates the di culty, since the logical structure of the theory
is generally reckoned to be represented by Dirac’s transformation
theory. According to it, any quantum state can indeed be regarded as
made up of other states – branches in an Everett-type ‘many-worlds’
picture. The di culty is that this representation is not unique. There
are many di erent ways in which one and the same state, formed
from the same two ‘observer’ and ‘object’ systems, can be
represented as being made up of other states. We can, for example,
use position states, but we can equally well use momentum states.
A DUALISTIC PICTURE
The purists among the quantum ‘founding fathers’, above all Dirac
and Heisenberg, saw a close parallel between the representation of
one and the same quantum state in many ways and the possibility of
putting many di erent coordinate systems on one and the same
space-time. In relativity, this corresponds to splitting space-time into
space and time in di erent ways. After Einstein’s great triumph, no
physicist would dream of saying that this could be done in one way
only. Similarly, Dirac and Heisenberg argued, there is nothing in
quantum theory to suggest that there is a preferred way to represent
quantum states. However, the parallel may not be accurate.
The italics are called for. We have reached the critical point. The
suggestion is that the universe as a whole is described by a single,
stationary, indeed static state. Why should this – with its implication
that nothing happens – be so? This is where we start to make
contact with the earlier part of the book. Time and change come to
an end when Machian classical dynamics meets quantum mechanics.
We have seen that a Machian universe should have only one value
of the energy: zero. We also know (Box 2) that a quantum theory
can be obtained by quantizing a corresponding classical theory. In
fact, it is easy to show that whereas quantizing Newtonian
dynamics, with its external framework of space and time, leads to
the time-dependent Schrödinger equation, quantizing the simple
Machian model considered in Chapter 7 leads to a quantum theory
in which the basic equation is not the time-dependent but the
stationary Schrödinger equation.
Before I can explain how this can be achieved, I must tell you what
the Schrödinger equation is like and what it can do. I believe it is
even more remarkable than physicists realize. This is where – if I am
right – we are getting near the secret of creation.
(Planck’s constant also occurs in the rst number, to ensure that all
three numbers have the same physical nature.)
Schrödinger won the 1933 Nobel Prize for Physics mainly for his
wave-mechanical calculation for the hydrogen atom. He found that
the energy eigenvalues of its stationary states are precisely the
energies of the allowed states in Bohr’s model. This was a huge
advance, since Schrödinger’s formalism had an inner unity and
consistency to it completely lacking in the older model. Brilliant
successes of the new wave mechanics, many achieved by
Schrödinger himself, soon came ooding in, leaving no doubt about
the great fruitfulness of the new scheme.
We must now see if we can dispense not only with time but also
with absolute space in quantum mechanics. In a timeless system the
energy E is zero, and the condition in Box 13 says simply that at
every point of Q the sum of the curvature number and the potential
number is zero. The potential number is already in the form we
need. For any possible relative con guration, the potential has a
unique value: it depends on nothing else. To nd the potential
number, we simply calculate the potential V for each con guration
and then multiply by , getting This part of the calculations is
pleasingly self-contained because V depends only on the relative
con guration. Each structure has its own potential irrespective of
how we imagine the structure to be embedded in space.
In fact, it is quite easy to see that the wave functions that satisfy
the Schrödinger conditions in this Machian case are precisely the
eigenfunctions of ordinary quantum mechanics for which the
angular-momentum eigenvalues are zero. This exactly matches our
result in classical mechanics – that the best-matching condition
leads to solutions identical to the Newtonian solutions with angular
momentum zero. We have already seen why they must be static
solutions.
The year 1980 was another turning point in my life. It was when
Bruno Bertotti and I thought we might have found a new theory of
gravitation, only to learn that the two ideas on which we had based
it were already an integral part of Einstein’s theory. Karel Kuchař’s
intervention rounded o our work but also brought it to an end. It
was something of an anticlimax. Bruno became increasingly
involved in experiments using spacecraft, aimed at detecting the
gravitational waves predicted by Einstein’s theory. For a year or two
I actually stopped doing physics and became politically active in the
newly founded Social Democratic Party (the SDP). However, the old
interests soon revived. Margaret Thatcher’s decisive general election
victory in 1983 hastened the process.
I rst met Lee a few weeks before I travelled to Salt Lake City in
the autumn of 1980. It was quite a dramatic time for me since I had
just narrowly escaped death through an insidious appendix that had
burst without giving me any pain. My only symptoms were
tiredness, slight sickness and the merest hint of stomach pain.
Luckily my vigilant doctor sent me to hospital as a precaution. An X-
ray proved di cult to interpret, and after quite lengthy deliberation
the doctors decided to open me up. They found that any further
delay could have been fatal. Seeing my state, the surgeon apparently
commented that ‘this must be a very brave man’, believing I must
have been in agony. In fact, I had been cheerfully reading The Times
without any discomfort only half an hour before the operation. The
day after I came back from hospital still convalescing, two American
physicists visiting Oxford phoned to say that they had heard from
Roger Penrose about my interest in Mach’s principle. Could they
come and see me? They came the next day, and I greeted them in
my dressing gown.
One was Lee, then a young postdoc. The meeting changed both
of our lives signi cantly. He proved very receptive to the ideas of
Leibniz and Mach to which I introduced him, while he encouraged
me to see what application they might have to the problem to which
he had decided to devote himself – quantum gravity. We met several
times in the next few years, and collaborated on an attempt to
formulate Leibniz’s philosophical system, his ‘monadology’, in
mathematical form. I think we made some real progress. Lee has
written about his view of things in his The Life of the Cosmos. Certain
aspects of our work together were decisive in my own elaboration of
the notion of time capsules and my conviction that the ultimate and
only truly real things are the instants of time. As far as I am aware,
Leibnizian ideas o er the only genuine alternative to Cartesian-
Newtonian materialism which is capable of expression in
mathematical form. What especially attracts me to them is the
importance, indeed primary status, given to structure and
distinguishing attributes, and the insistence that the world does not
consist of in nitely many essentially identical things – atoms
moving in space – but is in reality a collection of in nitely many
things, each constructed according to a common principle yet all
di erent from one another. Space and time emerge from the way in
which these ultimate entities mirror each other. I feel sure that this
idea has the potential to turn physics inside out – to make the
interestingly structured appear probable rather than improbable.
Before he became a poet, T. S. Eliot studied philosophy. He
remarked, ‘In Leibniz there are possibilities.’
A SIMPLE-MINDED APPROACH
You can play di erent games in one and the same arena. You can
also adjust the rules of a game as played in one arena so that it can
be played in a di erent arena. Both general relativity and quantum
mechanics are complex and highly developed theories. In the forms
in which they were originally put forward, they seem to be
incompatible. What I found to my surprise was that it does seem to
be possible to marry the two in Platonia. The structures of both
theories, stripped of their inessentials, mesh. What if Schrödinger,
immediately after he had created wave mechanics, had returned to
his Machian paper of only a year earlier and asked himself how
Machian wave mechanics should be formulated? His Machian paper
implicitly required Platonia to be the arena of the universe, while
any wave mechanics simply had to be formulated on a con guration
space. Such is Platonia, though it is not quite the hybrid Newtonian
Q he had used. But the structure of Machian wave mechanics would
surely have been immediately obvious to him, especially if he had
taken to heart Mach’s comments on time. As a summary of the
previous chapter, here are the steps to Machian wave mechanics in
their inevitable simplicity.
But if the N particles are the complete universe, there cannot be any
variation with change of centre of mass, orientation or time for the
simple reason that these things do not exist. The Machian wave
function of the universe has to be simply
Note the grander ψ. This is the wave function of the universe. It has
found its home in Platonia.
From 1955 to about 1970, much work was done along these
lines in studies of a space-time which is almost at and therefore
very like Minkowski space (I did my own Ph.D. in this eld). In this
case, the parallel between Einstein’s gravitational eld and
Maxwell’s electromagnetic eld becomes very close, and a
moderately successful theory (experimental veri cation is at present
out of the question, gravity being so weak) was constructed for it.
Within this theory it is certainly possible to talk about gravitons;
like photons, they have only two degrees of freedom. However,
Dirac and ADM had set their sights on a signi cantly more
ambitious goal – a quantum theory of gravity valid in all cases. Here
things did not match up. The expected two true degrees of freedom
did not tally with the three found from the analysis of general
relativity as a dynamical theory – as geometrodynamics.
The problem was that no clear choice could be made. Any and
all 3-spaces can appear in the relations that summarize so
beautifully the true essence of general relativity. What is more, any
choice would ultimately amount to the introduction of distinguished
coordinates on space-time. But this would run counter to the whole
spirit of relativity theory, the essence of which was seen to be the
complete equivalence of all coordinates. So if a choice were made,
the price would be the loss of this equivalence. The price and the
problem are one and the same. They presented the quantum
theoreticians with a head-on collision between the basic principles
of their two most fundamental theories – the need for a de nite
time in quantum mechanics and the denial of a de nite time in
general relativity. At an international meeting on quantum gravity
held at Oxford in 1980, Karel Kuchař, concluding his review of the
subject, stated that the problem of ‘quantum geometrodynamics is
not a technical one, but a conceptual one. It consists in the
diametrically opposite ways in which relativity and quantum mechanics
view the concept of time’. I have added the italics. I was there to hear
the talk, and Kuchaf’s comment made a deep impression on me.
The search for the third of space that would become time has
been like The Hunting of the Snark, Lewis Carroll’s mythical beast
that no one could nd. Since the idea of intrinsic time was rst
clearly formulated about thirty- ve years ago, the beast has not
been found. Karel has done more than anyone else to try to track it
down. If he cannot nd it, I feel that comes quite close to a non-
existence proof. My own belief is that the idea is based on an
incorrect notion of time. It is a mythical beast invoked in vain to
solve a titanic struggle. It does not surprise me that a special time
has not been found lurking in the tapestry of space-time. All I see in
that tapestry are change and di erences – and the di erences are
measured democratically. The idea of a special intrinsic time to be
extracted out of space, or out of any part of space-time or its
contents, violates the democratic theory of emphemeris time that
lies at the heart of general relativity.
If we look at the Newtonian parallel of the notion, it seems
strange. In a world of three particles, it is like saying that one of the
sides of the triangle they form is time while the other two are true
degrees of freedom. Such an attempt to nd time breaks up the
unity of the universe. No astronomer observing a triple-star system
would begin to think like that. The key property of astronomical
ephemeris time is that all change contributes to the measure of
duration. There has to be a di erent way to think about time.
DeWitt already clearly saw the problem posed at the end of the last
chapter – the crass contradiction between a static quantum universe
and our direct experience of time and motion – and hinted at its
solution in 1967. Quantum correlations must do the job. Somehow
they must bring the world alive. I shall not go into the details of
DeWitt’s arguments, since he saw them only as a rst step. However,
the key idea of all that follows is contained in his paper. It is that the
static probability density obtained by solving the stationary
Schrödinger equation for one xed energy can exhibit the
correlations expected in a world that does evolve – classically or
quantum mechanically – in time. We can have the appearance of
dynamics without any actual dynamics.
The key step now is to divide the total pattern of each wave into
a regular part, corresponding to an imagined perfectly sinusoidal
behaviour, and a remainder that is the di erence between it and the
actual (nearly sinusoidal) behaviour. Call this the di erence pattern
(there is one for each mist). If the condition of approximate phase
locking holds, it turns out that the di erence patterns satisfy with
respect to our ‘space’ and ‘time’ an equation of the same form as the
time-dependent Schrödinger equation, except for the appearance of
one additional term. This term will have less and less importance,
though, the more closely the assumptions of the semiclassical
approach are satis ed.
In fact, the way in which a wave packet will move is coded in the
relative positioning of the crests and troughs of the red and green
mists. We see this most clearly in momentum eigenstates. If the red
crests are ahead of the green crests, they go one way, but if the crest
positioning is reversed, they go the other way.
One thing is clear: the origin of our belief that the universe is
expanding cannot be coded in the relative positioning of the crests of
the two waves, for the designations ‘red’ and ‘green’ are purely
conventional. The ‘colours’ could be swapped, and nothing
observable would change. The argument that mere static positioning
of crests can correspond to what we call expansion of the universe is
a chimera. This was clearly recognized in 1986 by my German
physicist friend Dieter Zeh, who commented that it has meaning only
if an absolute time exists. It really is necessary to think very
di erently about these things if time is abolished once and for all as
an independent element of reality.
THE IDEA OF TIME CAPSULES: THE KINGFISHER
From 1988 to 1991 I was absorbed by this issue. I became more and
more convinced that a decisive new idea was needed, but for a long
time could nd no answer that satis ed me. I formulated the
problem this way. I imagined myself watching some phenomena
involving motion in a very essential and vital way – a display of
acrobatics, say, or the ight of a king sher. I then imagined being
struck dead instantaneously and my ‘soul’ being carried down to a
kind of Plato’s cave. Here I would nd omniscient mathematicians
examining a model of Platonia all covered with these red, green and
blue quantum mists that I have asked you to conjure up in your
mind’s eye. They are examining the solution of the Wheeler-DeWitt
equation corresponding to the universe in which I had just been
taken from life. I then asked myself this: what precise thing in that
mysterious pattern of mists blanketing Platonia corresponds to my
being aware of seeing the king sher in ight? Where – in a timeless
static world – is the appearance of motion coded? Where can I see
the king sher’s colours ashing in the sunlight?
I do not think there can be any. But there can be something else.
As I mentioned in Part 1, nobody really knows what it is in our
brains that corresponds to conscious experience. I make no pretence
to any expertise here, but it is well known that much processing goes
on in the brain and, employing normal temporal language, we can
con dently assert that what we seem to experience in one instant is
the product of the processing of data coming from a nite span of
time.
Now, at all the corresponding points the blue mist will have a
certain intensity, for in principle the laws of quantum mechanics
allow the mist to seep into all the nooks and crannies of Platonia.
Indeed, the rst quantum commandment is that all possibilities must
be explored. But the laws that mandate exploration also say that the
blue mist will be very unevenly distributed. In some places it will be
so faint as to be almost invisible, even with the acuity of vision we
acquire in Plato’s cave for things mathematical. There will also be
points where it shines with the steely blue brilliance of Sirius – or the
king sher’s wings. And again my conjecture is this: the blue mist is
concentrated and particularly intense at the precise point in Platonia
in which my brain does contain those perfectly coordinated
‘snapshots’ of the king sher and I am conscious of seeing the bird in
ight.
In fact, work that Hamilton did about ten years after his optical
discoveries shows how apt such a ‘Wheelerism’ is. As we saw in Part
2, classical physics is the story of paths in con guration spaces. They
are Newtonian histories. Hamilton thought about what would
happen if for them only one value of the energy is allowed, and
made a remarkable discovery. He found that just as light rays, which
are paths, arise from the wave theory of light when there is a regular
wave pattern, the paths of Newtonian dynamical systems can arise in
a similar fashion. I need to spell this out.
Working entirely within the framework of Newtonian dynamics,
Hamilton introduced something he called the principal function. All
you need to know about this function is that it is like the mists on
con guration space: at each point of the con guration space, it has a
value (intensity), the variation of which is governed by a de nite
equation. Hamilton showed that when, as can happen, the intensity
forms a regular wave pattern, the family of paths that run at right
angles to its crests are Newtonian histories which all have the same
energy. They are not all the histories that have that energy, but they
are a large family of them. Each regular wave pattern gives rise to a
di erent family. Hamilton also found that the equation that governs
the disposition of the wave crests, which in turn determine the
Newtonian histories, has the same basic form as the analogous
eikonal equation in optics. But whereas that equation operates in
ordinary three-dimensional space, this new equation operates in a
multidimensional con guration space.
I feel sure that the mystery of our deep sense and awareness of
history can be unravelled from the timeless mists of Platonia through
the latent histories that Hamilton showed can be there. But just how
is the connection to be made? In the remainder of this chapter I shall
explain Schrödinger’s valiant, illuminating, but unsuccessful attempt
to manufacture a unique history out of Hamilton’s many latent
histories. Then, in the next chapter, I shall consider the alternative –
that all histories are present.
The quantum measurement laws now tell us that one and only
one of the atoms will be ionized. It is selected by pure chance – it
can be anywhere in the spot. Once again, the entire wave function
that ‘bathes’ the other atoms is instantly destroyed, and a new
narrow beam continues outward from the second ionized atom. The
same process of ionization, collapse and ‘jet formation’ is repeated at
each successive shell. For an alpha particle with su cient energy,
this may happen hundreds or even thousands of times. A track is
formed. It has some important features.
The central question of this fth part of the book is this: whence
history?
What light does Bell’s rst account cast on this question? What
are the essential elements that go into the creation of history? Bell’s
analysis promises to give us real answers to these questions, since an
alpha-particle track can truly be seen as prototypical history. All the
elements are there – a unique succession of events, a coherent story
and qualitative change as it progresses. It even models birth – when
the particle escapes from the radium atom – and demise – when it
nally comes to rest. It literally staggers to its death. The laws that
govern the unfolding of history are beautifully transparent. They
combine, in an intriguing way, causal development – the forward
thrust of the track – with unpredictable twists and turns governed
only by probability. History is created by what looks like a curious
mixture of classical and quantum mechanics – the continuous track
and the twists and turns, respectively.
Before taking the next step, jettisoning collapse, we can add some
re nements. In the collapse picture, we can not only mark (with
‘paint’) the con guration point that is the time capsule of the
complete track. We can imagine a snapshot taken when only, say,
557 atoms have been ionized. The con guration point captured by it
will also be a time capsule, and we can mark it too. If we mark in
this manner all the stages – from no ionizations to all ionizations –
all the corresponding time capsules will be di erent points in the
con guration space. That is because they tell di erent stories, some
of which only reach, say, the track’s ‘adolescence’ or ‘middle age’.
Di erent con guration points necessarily represent di erent stories.
However, they are joined up more or less continuously in a path,
which represents an unfolding process.
If, like the god I imagined come to look at Platonia and its mists,
we could ‘see’ the con guration space and the wave function
sweeping over it, then in Bell’s ‘crude’ account we should see a patch
of wave function jigging its way along a track. The points along it
are the complete cloudchamber con gurations with successively
more ionizations. This con guration track is quite unlike the track
that represents a history in Newtonian dynamics. For a single alpha
particle, that is a track in three-dimensional space and the points
along it, de ned by three numbers, cannot possibly record history. In
contrast, each of the points traced out in the big con guration space
looks like a history of the three-dimensional track up to some point
along it. An analogy may help. Doting parents take daily snapshots of
their child and stick them day by day into a progress book. The
progress book after each successive day is like each successive point
along the track in the big con guration space: it is the complete
history of the child up to that date. Similarly, a point along the track
does not show the alpha particle at an instant of time, but its history
up to that time.
It is not easy to explain why it behaves like this, but let me try.
The most important thing is that a con guration space is not some
blank open space like Newton’s absolute space, but a kind of
landscape with a rich topography. Think of the wave function
pouring forth like oodwater sweeping over a rocky terrain, whose
features de ect the water. It will help if you look again at Triangle
Land (Figures 3 and 4). It is bounded by sheets and ribs, and is the
con guration space for just three particles. The con guration space
for 1027 particles is immensely more complicated. Things like the
ribs and sheets that appear as boundaries of Triangle Land occur as
internal topography in Platonia, which is traversed by all kinds of
structures. The rules that govern the evolution of the wave function
force it to respond to this rich topography. The wave-function
laments are directed by salient features in the landscape.
Now that we have some idea of how the ‘ rework explodes’, we
can think about its interpretation. The problem is that we never see
con guration space. That is a ‘God’s-eye’ view denied to our senses –
but fortunately not to our imaginations. We also never see a solitary
alpha particle making many tracks at once: all we ever see is one
track. How is this accounted for in the second scenario? By the same
device as before – by collapse. In the rst scenario, the alpha particle
was in many di erent places in its con guration space
simultaneously before we forced it to show itself in one region. This
was done by making it interact with an atom. This, most
mysteriously, triggered collapse, which was repeated again and
again.
The reason for this is that seeds of the many di erent tracks –
di erent histories – are already contained in the initial wave
function. A concentrated wave function necessarily spreads, and if
this happens in a large enough con guration space under low-
entropy conditions it can excite many di erent con gurations that
embody records of many di erent histories. There is a snowball
e ect. We start with many small snowballs, the di erent possibilities
for the alpha particle at the beginning of the process. Each possibility
then becomes associated – entangled – with a di erent track. This is
rather like many di erent snowballs picking up snow. Subject always
to a pervasive quantum uncertainty, a fuzziness at the edges, these
are Everett’s many worlds. The distinctness of these di erent worlds,
the di erent histories, is determined by the extent to which part of
the system (the alpha particle in this case) is in the semiclassical
(geometrical-optics) regime.
So this is the next twist in the saga. First Hamilton found families
of classical, particle-like histories as ‘light rays’ in a regular
(semiclassical) wave eld. Then Schrödinger tried to mimic particle
tracks by superposing many slightly di erent semiclassical solutions
to create just one wave packet – the model of a single particle. It was
rather hard and contrived work for a meagre – but still very beautiful
– result. However, it immediately slipped through his ngers. But
then Heisenberg and Mott showed that quantum mechanics could
work far more e ectively as the creator of history than Schrödinger
had ever dreamed. Now one single semiclassical solution generates
(before the nal collapse) many histories. Instead of Schrödinger’s
contrived
The story goes on. We have put only the cloud chamber into the
quantum mill – can we put the universe, ourselves included, in too?
That will require us to contemplate the ultimate con guration
space, the universe’s.
You can surely see where this is leading. Now the snowballs can
grow to include us and our conscious minds, each in di erent
incarnations. They must be di erent, because they see di erent
tracks; that makes them di erent. These similar incarnations seeing
di erent things necessarily belong to di erent points in the
universal con guration space. The pyrotechnics of wave-function
explosion out of a small region of Platonia – the decay of one
radioactive nucleus – has sprinkled ery droplets of wave function
at precise locations all over the landscape. (What an awful mixing of
metaphors – snowballs and sparks! But perhaps they may be
allowed to survive editing. The snowballs are in the con guration
space, the sparks in the wave function. This is a dualistic picture.)
And now to the great Everettian di erence: collapse is no longer
necessary. Nothing collapses at all. What we took to be collapse is
more like waking up in the morning and nding that the sun is
shining. But it could have been cloudy, or cloudy and raining, or
clear and frosty, or blowing a howling gale, or even literally raining
cats and dogs. When we lay down to sleep in bed – when we set up
the alpha-particle experiment – we knew not what we should wake
to. What we take to be wave-function collapse is merely nding that
this ine able self-sentient something that we call ourselves is in one
point of the con guration space rather than another. When we
observe the outcome of an experiment, we are not watching things
unfold in three-dimensional space. Something quite di erent is
happening. We are nding ourselves to be at one place in the
universal con guration space rather than another. All observation,
which is simultaneously the experiencing of one instant of time, is
ultimately a (partial) locating of ourselves in Platonia. Each of our
instants is a self-sentient part of a Platonic form.
Bell claimed that the really novel element in Everett’s theory had
not been identi ed. This was ‘a repudiation of the concept of the
“past”, which could be considered in the same liberating tradition as
Einstein’s repudiation of absolute simultaneity’. Obviously,
something exciting is in prospect, and Bell does not disappoint. He
looked for the quantum property that enabled Everett to make his
many-worlds idea plausible, and pointed out that the accumulation
of mutually consistent records is a vital part of it. This recognition
had led Bell to his analysis of the formation of alpha-particle tracks,
which have the obvious interpretation that they are records of
alpha-particle motion. He showed that ‘record formation’ is a
characteristic quantum property. At least under cloud-chamber
conditions, the wave function concentrates itself at con guration
points that can be called records. Although Bell did not use my
term, such points are manifestly time capsules. He noted that
Everett’s interpretation could not even be formulated were it not for
the wave function’s propensity to nd them.
This is all very entertaining – and I too have children and life
insurance – but these are just the kind of ad hominem quips that
were tossed at Copernicus and Galileo. I do believe that Bell came
close to a viable cosmological interpretation of quantum mechanics,
and should have kept faith with his title (‘Quantum mechanics for
cosmologists’). But he left the cosmologists with nothing. Later he
gave warm support to one of the theories in which wave-function
collapse is a real physical process. In it, the propensity of the
quantum-mechanical wave function to nd time capsules plays no
role. History is created by a succession of actually realized states. It
is there with or without any record of it.
From the way Bell wrote in 1980, either he was unaware of the
Wheeler-DeWitt equation and the possibility that the universal wave
function is static, or he dismissed this without mention. It would be
interesting to know how he would have reacted to the idea – he
seems to have had a somewhat Newtonian notion of time. Sadly, he
died several years ago, so we cannot ask him. I regret this especially
since his 1980 proposal is very close to mine in two of its three
main elements. He may have believed in time, but his emphasis on
memories and records and their rather natural occurrence in the
quantum context are valuable support for me. So are his views on
ontology and psychophysical parallelism. This is the third common
element.
In discussing Everett’s theory, I mentioned the so-called
preferred-basis problem. This arises from transformation theory: a
quantum state simultaneously encodes information about mutually
exclusive properties. Viewed one way, it gives probabilities for
particle positions; viewed another, it gives probabilities for their
momenta. It is impossible to extract this information simultaneously
and directly by, so to speak, ‘looking at the system’. We must let the
system interact with instruments. Depending on how the
instruments are arranged, we can extract information about either
the positions or the momenta, but not both at once. The ambiguity
becomes especially acute if the instruments are treated quantum
mechanically. We cannot say what state they are in or what they are
measuring.
This book has been one long, sustained e ort to shed redundant
concepts. We now are down to two: a static but well-behaved wave
function and the con guration space. The latter is Platonia, our
pitch. I look at it as a child might – what a lopsided thing it is!
However I turn in my mind the notion of ‘thing’, the space of all
things constructed according to one rule comes out asymmetric. All
the mathematical structures built by physicists to model the world
have this inherent asymmetry. One rule creates triangles, but they
are all di erent. No matter how you arrange them, their
con guration space falls out oddly. Have another look at Triangle
Land (Figures 3 and 4), which is just about the simplest Platonia
there can be. And what does it look like? An upturned Matterhorn.
Imagine trying to play football on that pitch.
But ‘can’ is not ‘must’. The fact is that Mott used a special
technique, always followed in such calculations, that mimics the
wave-packet behaviour. The answer is to some extent simply
assumed rather than truly derived and demonstrated. This can be
done because the time-dependent and stationary Schrödinger
equations have di erent structures, the latter having an extra
freedom not present in the former. At each stage of his calculations,
Mott systematically exploited this extra degree by making a de nite
kind of choice. This choice was not imposed by the mathematics but
was made, probably instinctively, to match his temporal intuition.
In fact, Mott’s solution is not a proper solution at all but a kind of
bookkeeping record of how the real process would unfold in time. In
addition, the condition corresponding to low entropy was also
assumed rather than derived.
One con guration at which the blue mist does shine brightly will
be a characteristic distribution of all the particles in the Sun. To an
experienced astrophysicist, this distribution tells an immensely rich
story stretching back to the rst three minutes (in the standard
picture) when the primordial hydrogen and helium abundances
were established. The whole story of the cosmos that we call our
own is written in the distribution of the Sun’s particles: the
formation of galaxies and the earliest generations of stars; the
supernova explosion that triggered the formation of the Sun and the
solar system, and left the radioactivity that still powers so much
tectonic and volcanic activity on the Earth; and the Sun’s steady
burning of its nuclear fuel.
No one has done more than Roger Penrose to highlight this fact.
His The Emperor’s New Mind has an entertaining illustration of the
creating divinity seeking with a pin to nd the tiny improbable
point of the initial condition of the universe from which its utterly
unlikely history must have sprung. Penrose seeks to explain this in a
theory in which both time and actual quantum-mechanical collapse
are real, and the laws of nature are inherently asymmetric in time.
My approach is quite di erent because I think that the whole
problem of time and its arrow can – paradoxically – be formulated
more precisely and transparently in a context in which time does
not exist at all. I also believe that, far from being highly unlikely,
the kind of history and cosmos we experience are characteristic and
likely in a timeless scenario.
Moreover, the paths are still only ‘seeds’. The nding of the full
rich structures which tell us so insistently that time exists and ows
must result from entanglement with the host of the remaining
quantum variables that constitute the expanses of Platonia. When
discussing alpha-particle tracks, I emphasized that Mott employed a
special device to concentrate the wave function on time capsules.
Considered purely in terms of the stationary Schrödinger equation,
this was arti cial. This is what created the static alpha-particle
tracks and such a strong sense of time and history out of the ‘seed’
of a spherical wave pattern.
A WELL-ORDERED COSMOS?
Let me end the main part of the book with a few comments on
structure, and what strikes a theoretical physicist as improbable. If
we think that dynamical histories in space and time are the
fundamental things in nature, then all statistical re ections on the
world lead to great di culties. Most histories are unutterably boring
over all but a minuscule fraction of their length. We can never
understand the miracle of the structured world. Things are
completely changed if quantum cosmology is really about some
well-behaved distribution of a static wave function over Platonia.
The con gurations at which ψ collects strongly must be special – in
some sense they must resonate with all the other con gurations that
are competing for wave function. Quantum cosmology becomes a
kind of beauty contest in eternity. The winners – those that get a
high probability density – must be exceptional, like the DNA
molecule. This is just the opposite of what classical physics leads us
to expect. There, the winners are boring.
The single most striking thing about the universe we see around
us is its rich structure, which is so di cult to understand on a priori
statistical grounds. Until the modern scienti c age, all thinkers saw
the rst task of science as being the direct description and
explanation of this structure. This natural impulse is re ected in the
Pythagorean notion of the well-ordered cosmos. It was still very
strong in both Kepler and Galileo. However, when Newton
demonstrated the supreme importance of accelerations in dynamics,
the perspective of science changed, for the world at the present
instant became the mere consequence of its initial conditions.
Instead of asking directly how structure is fashioned, science turned
to asking how it is refashioned.
Pied Beauty
The evidence for them is strong. The history of science shows that
physicists have tended to be wrong when they have not believed
counterintuitive results of good theories. However, despite strong
intellectual acceptance of many worlds, I live my life as if it were
unique. You might call me a somewhat apologetic ‘many-worlder’!
There are occasions when the real existence of other worlds, other
outcomes, seems very hard to accept. Soon after I started writing
this book, Princess Diana was killed, and Britain – like much of the
world – was gripped by a most extraordinary mood. Watching the
funeral service live, I did wonder how seriouslyone can take a
theory which suggests that she survived the crash in other worlds.
Death appears so nal.
The same applies to us, for our conscious instants are embedded
in the Nows. The probability of us experiencing ourselves doing
something is just the sum of the probabilities for all the di erent
Nows in which that experience is embedded. Everything we
experience is brought into existence by being what it is. Our very
nature determines whether we shall or shall not be. I nd that
consoling. We are because of what we are. Our existence is
determined by the way we relate to (or resonate with) everything
else that can be. Although Darwinism is a marvellous theory, and I
greatly admire and respect Richard Dawkins’s writings, one day the
theory of evolution will be subsumed in a greater scheme, just as
Newtonian mechanics was subsumed in relativity without in any
way ceasing to be great and valid science. For this reason, and for
the remarks just made, I do not think that we are robots or that
anything happens by chance. That view arises because we do not
have a large enough perspective on things. We are the answers to
the question of what can be maximally sensitive to the totality of
what is possible. That is quite Darwinian. Species, ultimately genes,
exist only if they t in an environment. Platonia is the ultimate
environment.
You will naturally ask why we do not hear this music of the
spheres. Keats provides a rst answer: ‘Heard melodies are sweet,
but those unheard Are sweeter’. But Leibniz may have given the true
answer. In his monadology, he teaches that the quintessential you,
everything you experience in consciousness and the unconscious, is
precisely this music. You are the music of the spheres heard from
the particular vantage point that is you. This is taking a little liberty
with the letter but certainly not the spirit of his great philosophical
scheme. On the subject of liberties, I have taken fewer with Leibniz
than Michael did with Shakespeare. Hal does not actually ask
Falsta (‘fat-witted with drinking of old sack’) why he should be so
super uous ‘to inquire the nature of time’ but to ‘demand the time
of day’. But, were it not for the blessed Sun and its diurnal rotation
(our fortunate circumstances), the one question would be as
profound as the other.
Where Is Heaven?
I have long thought that, if only we had the wit to see it, we are
already in heaven. It is Platonia. I say this with some trepidation,
though I believe it is true. If so, Platonia must be hell and purgatory
as well. What I mean by this is really quite simple: some places in
Platonia are very admirable, pleasant and beautiful, many are
boring in the extreme, and others are horrendously nasty. The same
contrasts exist within the individual Nows. What we do not know is
where the wave function collects.
I can only say that is not how I see things. I search in vain for
Omega in Platonia and nd only Alpha. But Platonia is a vast land.
Let us cherish everything around us wherever we happen to nd
ourselves in the Platonic palace.
Of one thing I feel very sure. Many poets and theologians give a
misleading image of heaven and eternity. Consider the opening lines
of Vaughan’s famous poem The World’:
I saw Eternity the other night
Like a great Ring of pure and endless light,
Doesn’t the Denial of Motion Take All Joy and Verve out of Life?
I do feel this issue keenly. The king sher parable should make that
clear. In principle, there is no reason why we should not attempt to
put our very direct sense of change directly into the foundations of
physics. There is a long tradition, going back at least to Hamilton,
that seeks to make process the most basic thing in the world.
Roughly, the idea is that physics should be built up using verbs, not
nouns. In 1929 the English philosopher Alfred North Whitehead
published an unreadable – in my experience – book called Process
and Reality in which he advocated process. It all sounds very
exciting, but I just do not think it can be done, despite a valiant
attempt by Abner Shimony. Having translated seventy million words
of Russian into English, I can say with some feeling that sentences
do have a subject and generally an object. I could have written this
book using the one verb ‘to be’, which hardly counts as a verb. For
this reason it seldom appears in Russian; when it does, it is most
often as a surrogate: ‘to appear’. But a book without nouns is
nothing. Not even James Joyce could write it. For some reason,
disembodied verbs exert a fascination not unlike the grin of the
Cheshire cat. But when Owen Glendower claimed to be able to ‘call
up spirits from the vasty deep’, Hotspur answered: ‘Aye, and so can I
and any man, but will they come when you call them?’ I should like
to see it done.
Keats too, for all the beauty of his Grecian urn, addresses it with the
words
When Keats wrote these lines, he must have known that all too soon
his home would be a grave. Is Platonia a graveyard? Of a kind it
undoubtedly is, but it is a heavenly vault. For it is more like a
miraculous store of paintings by artists representing the entire range
of abilities. The best pictures are those that somehow re ect one
another. These are the paintings we nd there in profusion. There
are very few of the mediocre, dull ones. Despite what Ladislaw says,
the best paintings have a tremendous vibrancy. Turner does almost
bind you to the mast of the Ariel. Indeed, Ladislaw’s own words
immediately before the passage quoted above are: ‘After all, the true
seeing is within.’ Frozen it may be, but Platonia is the demesne
where ‘Beauty is truth, truth beauty’ and the boughs cannot shed
their leaves ‘nor ever bid the Spring adieu’. With that perfect ode,
Keats did achieve the immortality for which he so desperately
longed.
In a ne essay entitled ‘The timeless world of a play’, Tennessee
Williams praises great sculpture because it
often follows the lines of the human body, yet the repose of
great sculpture suddenly transmutes those human lines to
something that has an absoluteness, a purity, a beauty,
which would not be possible in a living mobile form.
He argues that a play can achieve the same e ect, and so help us to
escape the ravages of time. ‘Whether or not we admit it to ourselves,
we are all haunted by a truly awful sense of impermanence.’ Again,
this is very beautiful writing, but has Williams failed to see the truth
all around us – the Platonic eternity we inhabit in each instant? Is it
blindness that drives him to seek eternity? Some people can pass a
cathedral without noticing it.
This does express the main ideas I have tried to get across in the
nal part of the book. Each experienced instant is a separate
creation (birth), the ever inaugural act of existence, brought to life
by the gathering of all times. The thrill that Janet Baker experiences
in each Now is the assolutamente unico ma imprevedibile presente, that
nding of ourselves in one of the instants that quantum mechanics
makes resonate especially strongly with other instants.
No, this is home. Mach once commented that ‘In wishing to preserve
our personal memories beyond death, we are behaving like the
astute Eskimo, who refused with thanks the gift of immortality
without his seals and walruses.’ I am not going without them, either.
I cannot even if I wanted to: they are part of me. Like you, I am
nothing and yet everything. I am nothing because there is no
personal canvas on which I am painted. I am everything because I
am the universe seen from the point, unforeseeable because it is
unique, that is me now. C’est moi. I am bound to stay. We all watch
—and participate in—the great spectacle. Immortality is here. Our
task is to recognize it. Some Nows are thrilling and beautiful beyond
description. Being in them is the supreme gift.
NOTES
PREFACE
(2) (p. 4) On hearing about my plans for this book, Michael Purser
brought to my attention the following rebuke from Prince Hal to
Falsta :
Among the popular books that I know, the two that undoubtedly
give most prominence to the problem of time in quantum gravity
are Lee Smolin’s The Life of the Cosmos, which contains some
discussion of my own ideas, and David Deutsch’s The Fabric of
Reality. There is considerable overlap between my book and
Deutsch’s chapter ‘Time: the rst quantum concept’. One technical
book, now going into a third edition, that from the start has taken
timelessness very seriously is Dieter Zeh’s The Physical Basis of the
Direction of Time.
It may be that the reason why a book like this one, devoted
exclusively to the idea that time does not exist, has not hitherto
been published by a physicist has a sociological explanation. For
professionals working in institutes and dependent on the opinions of
peers for research funding, such a book might damage their
reputation and put further research in jeopardy. After all, at rst it
does seem outrageous to suggest that time does not exist. It may not
be accidental that I, as an independent not reliant on conventional
funding, have been prepared to ‘come out’.
Do you believe time is a truly basic concept that must appear in the
foundations of any theory of the world, or is it an e ective concept
that can be derived from more primitive notions in the same way
that a notion of temperature can be recovered in statistical
mechanics?
Getting to Grips with Elusive Time (p. 17) The idea that instants
of time are distinct entities that should not be thought of as joined
up in a linear sequence is a powerful intuitive experience for at least
one non-scientist. A few days after the Sunday Times published its
article ‘Time’s assassin’ about my ideas in October 1998, I received
by email a ‘Question for Julian Barbour’ from Gretchen Mills
Kubasiak, who had read the article about me. She introduced herself
with: I am merely a girl who lives in Chicago, works for a
construction company and nds herself thoroughly captivated by
your ideas. In fact, I have been unable to think of little else this past
week.’ She asked if she could put a question to me. Well, who could
resist that request? I said yes, asking if by any chance, with her rst
name, she had German ancestry, and commented: ‘I guess you know
the German expression Gretchenfrage and its origin in Goethe’s Faust,
when Gretchen asks Faust about his attitude to religion and if he
believed in God. It was especially nice to get your Gretchenfrage.’
Subsequent correspondence persuades me that ‘merely a girl’ might
not be the most accurate description of her, since she is a voracious
reader and traveller (among much else). Some of her thoughts about
time are worth passing on:
Reading these comments again three months after they came, they
strike me as often very close to my position. Incidentally, I address
the original Gretchen’s questions (Glaubst du an Gott? Wie halt’s du es
mit der Religion?) in the Epilogue.
Note for physicists (p. 18): Space plays two roles in Newtonian
physics: it binds its contents together to form the plurality within
the unity mentioned in this section (the separations between N
objects in Euclidean space are constrained by both inequalities and
algebraic relations, which give expression to this unity) and if
de nes positions at non-coincident times. In the type of physics I am
advocating, only the rst property is used, as will become clear in
Part 3.
The Physical World and Consciousness (1) (p. 26) There is a clear
and detailed account of Boltzmann’s ideas in Huw Price’s book listed
in Further Reading.
(2) (p. 27) It is worth quoting here two passages from Boltzmann
himself. In 1895 he published (in perfect English—I wonder if he
had assistance) a paper in Nature with the title ‘On certain questions
of the theory of gases’. It ends with a truly remarkable and concise
statement of what much later became known as the anthropic
principle. This expression was coined in 1970 by the English
relativist Brandon Carter (who had earlier made important
discoveries about the physics of black holes in the period leading up
to Hawking’s discovery that they can evaporate). The anthropic
principle, which gained widespread attention initially through the
book The Anthropic Cosmological Principle by John Barrow and Frank
Tipler, expresses the idea that any universe in which intelligent life
exists must have special and unexpected (from a purely statistical
viewpoint) properties, since otherwise the intelligent life that
observes these properties could not exist. Therefore we should not
be surprised to nd ourselves in a universe that does have special
and remarkable properties.
We assume that the whole universe is, and rests for ever, in
thermal equilibrium. The probability that one (only one) part
of the universe is in a certain state, is the smaller the further
this state is from thermal equilibrium; but this probability is
greater, the greater the universe itself. If we assume the
universe great enough we can make the probability of one
relatively small part being in any given state (however far
from the state of thermal equilibrium), as great as we please.
We can also make the probability great that, though the whole
universe is in thermal equilibrium, our world is in its present
state. It may be sayd [sic] that the world is so far from thermal
equilibrium that we cannot imagine the improbability of such
a state. But can we imagine, on the other side, how small a
part of the whole universe this world is? Assuming the
universe great enough, the probability that such a small part of
it as our world should be in its present state, is no longer
small.
The Ultimate Arena (1) (p. 39) In this section I say that all
structures that represent possible instants of time are three-
dimensional. This is because the space we actually observe has three
dimensions. However, in some modern theories (super-string
theories) it is assumed that space actually has ten or even more
dimensions. All but three of the dimensions are ‘rolled up’ so tightly
that we cannot see them. In principle, my instants of time could t
into this picture. They would then have ten (or more) dimensions.
(1) (p. 61) I have written at considerable length about the early
history of astronomy and mechanics and the absolute versus relative
debate in my Absolute or Relative Motion? This has recently been
reprinted as a paperback with the new title The Discovery of
Dynamics (OUP, 2001). I still hope to complete a further volume
bringing the story up to the present, and much has already been
written, but my plans are in ux because of the developments
mentioned at the end of the Preface and at various places in these
notes. Readers wanting a full academic (and mathematical)
treatment of the topics presented in Parts 2 and 3 of this book are
asked to consult the above and the papers (Barbour 1994a, 1999,
2000, 2001), which cite earlier papers. For references to recent
developments see p. 358 and my website (www.julianbarbour.com).
(2) (p. 64) In the main body of the text, I mention the importance of
the fortunate circumstances of the world in enabling physicists to
avoid worrying about foundations. Another very important factor is
the clarity of the notion of empty space, developed so early by the
Greek mathematicians, which deeply impressed Newton. He felt that
he really could see space in his mind’s eye, and regarded it as being
rather like some in nite translucent block of glass. He and many
other mathematicians pictured its points as being like tiny identical
grains of sand that, close-packed, make up the block. But this is all
rather ghostly and mysterious. Unlike glass and tiny grains of sand,
which are just visible, space and its points are utterly invisible. This
is a suspect, unreal world.
We are not bound to hang onto old notions. We can open our
eyes to something new. Let me try to persuade you that points of
space are not what mathematicians would sometimes have us
believe. Imagine yourself in a magni cent mountain range, and that
someone asked, ‘Where are you?’ Would you kneel down with a
magnifying glass and look for that invisible ‘point’ at which you
happen to be in the ‘space’ that the mountain range occupies? You
would look in vain. Indeed, you would never do such a silly thing.
You would just look around you at the mountains. They tell you
where you are. The point you occupy in the world is de ned by what
the world looks like as seen by you: it is a snapshot of the world as
seen by you. Real points of space are not tiny grains of sand, they
are actual pictures. To see the point where you are in the world, you
must look not inward but outward.
The Inertial Clock (p. 99) Tait’s work, which I feel is very
important, passed almost completely unnoticed. This is probably
because two years later the young German Ludwig Lange introduced
an alternative construction for nding inertial frames of reference,
coining the expression ‘inertial system’. Lange deserves great credit
for bringing to the fore the issue of the determination of such
systems from purely relative data, but Tait’s construction is far more
illuminating. Lange’s work is discussed in detail in Barbour (1989)
and Tait’s in Barbour (forthcoming).
The Second Great Clock (p. 107) A very nice account of the history
of the introduction of ephemeris time was given by the American
astronomer Gerald Clemence (1957).
Platonia for Relativity (p. 167) This is a technical note about the
de nition of superspace. The equations of general relativity lead to
a great variety of di erent kinds of solution, including ones in
which there are so-called closed time-like loops. These are solutions
in which a kind of time travel seems to be possible. The question
then arises of whether a given solution of general relativity—that is,
a space-time that satis es Einstein’s equations—can be represented
as a path in superspace, in technical terms, as a unique succession of
Riemannian three-geometries. If this is always so, then superspace
does indeed seem a natural and appropriate concept. Unfortunately,
it is de nitely not so. There are two ways in which we can attempt
to get round this di culty. We could say that classical general
relativity is not the fundamental theory of the universe, since it is
not a quantum theory. This allows us to argue that superspace is the
appropriate quantum concept and that it will allow only certain
‘well-behaved’ solutions of general relativity to emerge as
approximate classical histories. For these, superspace will be an
appropriate concept. Alternatively, we could extend the de nition of
super-space to include not only proper Riemannian 3-geometries (in
which the geometry in small regions is always Euclidean), but also
pscudo-Riemannian 3-geometries (in which the local geometry has a
Minkowski type signature), and also geometries in which the
signature changes within the space. For the reasons given in the
long note starting on p. 348 below, I prefer the second option.
The above note was written before my new insights mentioned at the
end of the Preface. I now believe that there is a potentially much more
attractive resolution of the di culty: the true arena of the world is not
superspace but conformal superspace, which I describe on p. 350.
Catching Up with Einstein (1) (p. 175) Figure 30 is modelled
directly on well-known diagrams in Wheeler (1964) and Misner et
al. (1973).
Schrödinger’s Heroic Failure (p. 278) In the rst draft of this book
I included a long section on the very interesting interpretation of
quantum mechanics advanced originally by de Broglie, and revived
by Bohm, whose 1952 paper I strongly recommend to physicists
together with Peter Holland’s book (Holland 1993). With regret I
omitted it, as I felt that it made this book too long, especially since I
believe that the interpretation does not really solve the problem.
However, I particularly value the way in which it shows that all the
results of quantum mechanics can be obtained in a framework in
which positions are taken as basic. This made the theory attractive
to John Bell, as we shall see in the next chapters.
(2) (p. 309) I should emphasize that Mott, like Bell, never used any
expression like ‘time capsule’, and clearly did not think in such
terms about alpha-particle tracks. Neither did Mott’s work on alpha-
particle tracks seem to have prompted him to any intimations of a
many-worlds type interpretation of quantum mechanics. I learned
this from Jim Hartle. Over a decade ago, when collaborating with
Stephen Hawking in Cambridge, Jim lodged at his college, Gronville
and Caius (featured famously in Chariots of Fire), which was also
Mott’s college. Over dinner Jim asked Mott whether his paper had
not led him to anticipate some form of Everett’s idea, and was told
no. Apparently, all the ‘young Turks’ followed the Copenhagen line
without hesitation at that time. Shortly before his death about two
years ago, when he was still mentally very alert, I contact Mott and
asked if I could talk to him about his paper. Alas, he was too ill to
keep the appointment, telling his secretary he was very disappointed
‘since the man wanted to talk about work I did nearly sixty years
ago’.
DOWKER. I think that neither your version (which I’ll call JMWI)
nor BMWI allows us to make predictions about what we observe (so
I disagree with Everett’s statement ‘the theory itself predicts that our
experience will be what it in fact is’). Let me take your version.
There we have many con gurations at time t. The most serious
problem is that in a scheme like yours, in which all the possibilities
are realized, there is no role for the probabilities. The usual
probabilistic Copenhagen predictions for the results of our
observations cannot be recovered. An excellent reference which
analyses the MWI literature and the various attempts to derive the
Born interpretation from MWI is Adrian Kent [1990, International
Journal of Modern Physics, A5, 1745]. Adrian concludes that they
fail. I’ll just state again the main reason that they fail: when all the
elements in a sample space of possibilities are realized, then
probability is not involved. Your idea is that it is the sample space
itself, i.e. how many copies of each con guration are included in the
sample space, which is determined by the (squares of the)
coe cients of the terms in the wave function. That is all well and
good (if bizarre). But there’s no reason then to call those numbers
probabilities, and no way to recover the probabilistic predictions of
Copenhagen quantum mechanics. In fact the MWI proponents
themselves agree that the failure to reproduce the Copenhagen
predictions is a problem and do try to address it, but without
success.
The potential signi cance of the rst paper has already been
explained on p. 349/350. At this stage, I do not wish to make any
rm statements about this new work since it is incomplete and has
not yet been exposed to scrutiny by other physicists, but I can at
least give some idea of what is at stake. The basic issue is the status
of the relativity principle. When Einstein and Minkowski created
special relativity, they deliberately made no attempt to explain the
remarkable structure that their work had brought to light: the
existence of spacetime and its associated light cone, both being
re ected in the Lorentz invariance of the laws of nature. They
adopted Lorentz invariance, which assumes the existence of length,
as the basis of physics. In small regions of spacetime, this still
remains true in general relativity.
Taken together the two papers cited above suggest that all of the
presently known facts of relativity and electromagnetism can be
derived in a new and hitherto unsuspected manner from three
assumptions: 1) an independent time plays no role in dynamics; 2)
best matching (pp. 116/7) is the essential element in the action
principle of the universe; 3) any theory satisfying these principles
must have nontrivial solutions. It is the third assumption that makes
a dramatic di erence. Hitherto, in common with other colleagues, I
had assumed (see p. 181) many di erent theories could satisfy the
rst two conditions, but, as my collaborator Niall Ó Murchadha
discovered, this is not the case. The reasons for this and its
remarkable potential consequences are spelled out in the second
cited paper. It is frustrating not to be able to say more at the present
moment, but at a time of uncertainty about the nal outcome it is
better to say less rather than too much. My website
(julianbarbour.com) will carry more detailed information.
JB
January 2001
FURTHER READING
Coveney, Peter and High eld, Roger, 1991, The Arrow of Time,
Flamingo, London.
Davies, Paul, 1991, The Mind of God, Simon & Schuster, New York.
Einstein, Albert, 1960, Relativity: The Special and the General Theory.
A Popular Exposition, Routledge, London.
Greene, Brian, 1999, The Elegant Universe: Superstrings, Hidden
Dimensions, and the Quest for the Ultimate Theory, Vintage Books,
New York.
Guth, Alan H., 1997, The In ationary Universe: The Quest for a New
Theory of Cosmic Origins, Perseus Books Group, New York.
Rees, Martin, 1997, Before the Beginning: Our Universe and Others,
Simon & Schuster, London.
Rees, Martin, 1999, Just Six Numbers: The Deep Forces that Shape the
Universe, Weidenfeld & Nicolson, London.
Smolin, Lee, 1997, The Life of the Cosmos, Weidenfeld & Nicolson,
London (Oxford University Press, New York).
Thorne, Kip, 1994, Black Holes and Time Warps: Einstein’s Outrageous
Legacy, Norton, New York.
Weinberg, Steven, 1977, The First Three Minutes, Basic Books, New
York (André Deutsch, London).
Weinberg, Steven, 1993, Dreams of a Final Theory: The Search for the
Fundamental Laws of Nature, Vintage, London.
Will, Cli ord, 1986, Was Einstein Right?, Basic Books, New York.
BIBLIOGRAPHY
Brout, R., 1987, ‘On the concept of time and the origin of the
cosmological temperature’, Foundations of Physics, 17, 603.
Kiefer, C., 1997, ‘Does time exist at the most fundamental level?’, in
Time, Temporality, Now, H. Atmanspacher and E. Ruhnau (eds),
Springer, Berlin.
Pais, A., 1982, ‘Subtle is the Lord . . . ‘The Science and Life of Albert
Einstein, Oxford University Press, Oxford.
Pinker, S., 1997, How the Mind Works, Penguin, London/Norton,
New York.
Zeh, H.-D., 1992, The Physical Basis of the Direction of Time, 2nd edn,
Springer, Berlin (3rd edn, 1999).
INDEX
aberration 128
action 111–13
Alpha 42, 46
antiparlicles 191
atoms 24–5, 49
Augustine, St 11
Barbour, Julian
bicycle riding 88
bifurcation 55
bosons 191
brain
damage 33
motion processing 28–30, 267
organization 31–2
state 26
as a time capsule 32–3
Bruno, Giordano 75
calculus 63
Candide 111
causality 312–13
cells 33
Clarke, Samuel 63
clocks 93–108
de ned 135
earth’s rotation 106
gravitational e ects 154–5
inertial 99–104
solar 97, 98
solar system 106–7
stellar 97, 98
universe as 107–8
water-clock 95
complementarity 204
collinear 75
Confucius 17
Creator 326–7
degrees of freedom 79
Descartes, René
on existence 49–50, 53
on motion 61–2
on many-worlds 224–5
di erence
di raction 285
absolute 74–5
in space-time 147, 150, 151
dynamics 12
Earth
as a clock 106
rotation 32, 87
as a time capsule 33
eclipses 98–9
eigenfunctions 234
eigenvalues 234–5
electromagnetism 124–6
equation of time 98
faces 33
fermions 191
on time 2
fossils 33
gas 318
geometrodynamics 167
gravitons 244
Guichardet, A. 347
haemoglobin 48
heaven 327–8
Heraclitus 1, 330
Hipparchus 98
hyperplanes 143–4
I
identity of indiscernibles 85–6
in ation 359
interactions 104–5
self- 197
fringes 196
invariable plane 88
James, William 28
Lapchinskii, V. 258
Leibniz, Willhelm Gottfried 16, 63–4, 74, 110, 119, 240, 322
light
cone 148
di raction 285
interference 124, 125
polarization 244
quanta (photons) 187, 188, 191, 243–4
rays 269–73
refraction 110
wave theory 124, 125, 269–71
linearity
in quantum mechanics 225, 231, 275
of time 19, 27–8
Lovelock, James 4
on inertia 65–6
on time 67
many-instants 302–5
mass
centre of 79–81
of particles 243–4
mechanics 12
Megamolecule Land 43
Mercury 161
Middlemarch 330–1
angular 86–90
eigenstates 201
motion 61–70
on absolute motion 62
on absolute space 20, 62–3
on absolute time 20
on creation 22
on Descartes 62
laws of 12, 20–1, 22, 64
on making his discoveries 74
on motion 12, 20–1, 22, 62–3
Nows 16, 18, 34, 40–1, 43, 44–5, 55–6, 174, 177, 333
Einstein on 143
and experiences 51–2, 53–4
and memories 55–6
in relativity 142–6
self-awareness 52–3
observables 204
Omego 328
Ó Murchadha, Niall 35, 349, 352, 358
optics 269–71
optimization 109–13
Pangloss, Dr 111
paradigms 115
Parmenides 1
particles 185–92
mass of 243–4
Pavia 114
‘philosophical’ motion 61
physics 38
end of 13–14
quantum versus classical 13, 15–16
Plato 44, 45
Poincaré, Henri 71, 76, 128, 135–6, 152–3, 156, 180, 346
on duration 123
Ptolemy, Claudius 99
pulsars 166
Q space 209
realism 252
refraction 110
Reinsch, M 347
rest 126–7
Rubakov, V. 258
Saturn 89, 90
scalar elds 191
on ‘I’ 331–2
Nobel Prize for Physics 235
stationary equation 230, 231, 232–5, 253
time-dependent Schrödinger equation 230, 231, 258
wave-packets 275–80
self-awareness 52–3
self-interference 197
Shapere, A. 348
as a clock 106–7
space 138
space-like 149
specious present 28
speed 96, 97
group 278
St Augustine 11
‘state’ 26
strata 345
as a clock 97, 98
supersymmetry 192
syzgies 78
T
Tait, Peter 100–1
thermodynamics 23
3-spaces 171
time 17–19
de ned 31
time-independent (stationary)
time-like 148
unity of unities 75
universe 22
chronology of 313–15
as a clock 107–8
expansion 262–4
modelling 40–6
order of 23
wave function of 242
velocity 96, 97
group 278
Vilenkin, Alexander 352, 359
Voltaire 111
water-clock 95
website 5, 7, 344
Wheeler, John Archibald 38, 44, 136, 167, 169, 175–6, 246–7
X-rays 190
Zeno’s paradox 49