Você está na página 1de 324

TIME, CAUSALITY, AND THE QUANTUM THEORY

BOSTON STUDIES IN THE PHILOSOPHY OF SCIENCE

EDITED BY ROBERT S. COHEN AND MARX W. WARTOFSKY

VOLUME 19
HENR Y MEHLBERG
HENRY MEHLBERG

TIME, CAUSALITY,
AND THE
QUANTUM THEORY
Studies in the Philosophy of Science

VOLUME ONE
Essay on the Causal Theory of Time

Edited by

ROBERT S. COHEN
with a preface by
ADOLF GRUNBAUM

D. REIDEL PUBLISHING COMPANY

DORDRECHT: HOLLAND! BOSTON: U.S.A.


LONDON: ENGLAND
Library of Congress Cataloging in Publication Data

Mehlberg, Henry, 1904-1979


Time, causality, and the quantum theory.

(Boston studies in the philosophy of science; v. 19)


Bibliography: p.
Includes indexes.
CONTENTS: v.I. Essay on the causal theory of time. - v. 2. Time in
a quantized universe.
l. Space and time. 2. Causality (Physics). 3. Quantum theory. 4.
Physics-Philosophy. I. Title. II. Series.
Q174.B67 vol. 19 [QC173.59.S65] SOls [530.1 '1]
ISBN-I3: 978-90-277-1074-1 e-ISBN-I3: 978-94-009-8935-1
DOl: 10.1007/978-94-009-8935-1

The Essay in this volume is based on an unpublished translation, by Paul


Benacerraff, of Essai sur la theorie causale du temps, which has been revised
by Carolyn R. Fawcett and Robert S. Cohen.

Published by D. Reidel Publishing Company


P. O. Box 17, 3300 AA Dordrecht, Hoiland.

Sold and distributed in the U.S.A. and Canada


by Kluwer Boston Inc.,
190 Old Derby Street, Hingham, MA 02043, U.S.A.

In all other countries, sold and distributed by Kluwer Academic Publishers


Group, P.O. Box 322, 3300 AH Dordrecht, Holland

D. Reidel Publishing Company is a member of the Kluwer Group

All Rights Reserved


Copyright © 1980 by D. Reidel Publishing Company, Dordrecht, Holland
and copyrightholders as specified on appropriate pages within
No part of the material protected by this copyright notice may be reproduced or
utilized in any form or by any means, electronic or mechanical,
including photocopying, recording or by any informational storage and retrieval
system, without written permission from the copyright owner Softcover reprint of
the hardcover I st Edition 1980
This Work is dedicated
to

Karl R. Popper and


Eugene P. Wigner
T ABLE OF CONTENTS
PREFACE (Adolf Grunbaum) xiii

EDITORIAL NOTE xv

ACKNOWLEDGMENTS xix
INTRODUCTION TO VOLUMES I AND II

VOLUME I : ESSAY ON THE CAUSAL THEORY OF TIME

Part I: The Causal Theory of Time in the Works of

Its Principal Representatives


Introduction 39
I. Leibniz and the Beginnings of the Causal Theory of Time 42
1. The Relational Theory 42
2. The Causal Theory 45
II. Kant's Phenomenalist Interpretation of the Causal Theory of
Time 51
1. General Remarks on Kant's Phenomenalism 51
2. The Causal Theory of Simultaneity 54
3. The Causal Theory of Succession 57
4. Examination of Schopenhauer's Criticism 59
5. The Plurality of Concepts of Time in Kant 65
III. Lechalas' Adaptation of the Causal Theory of Time to the
Laws of pre-Einsteinian Physics 70
1. General Remarks on Lechalas' Theory 70
2. The Principle of Determinism, Considered as Defming the
Temporal Order 71
3. Physical Reversibility 78
4. Psycho-Physiological Reversibility 83
5. The Epistemological Interpretation of the Causal Theory 88
IV. The Relativistic Phase of the Causal Theory of Time: The
Axiomatic Systems of Robb and Carnap 91
1. Robb's System 91
2. Carnap's Systems 95
x TABLE OF CONTENTS
3. The Epistemological Priority of the Causal Relation 98

4. The Ontological Priority of the Causal Relation 101


V. The Relativistic Phase of the Causal Theory of Time: The
Work of Reichenbach 105
1. The Causal Theory 105
2. Criticism of the Principle of Marking 111
3. Criticism Continued: Subjective and Objective Time 113
4. Criticism, Continued and Concluded: Convention and
Reality in the Temporal Order 123
5. The Semi-Causal, Semi-Statistical Theory 125
6. The Branch Hypothesis 128
7. Conclusion 129
VI. Russell's Causal Explanation of Duration 134
1. The Logical Analysis of Duration 134
2. Russell's Causal Defmition 137
3. Criticism and Comment 138
4. Epistemological Remarks 141
5. Conclusion 145
VII. Alternative Approaches to Time's Arrow 147
Introductory 147
1. K. G6del and S. Chandrasekhar 148
2. G. J. Whitrow 152
3. A. Grtinbaum and O. Costa de Beauregard 153
4. R. Schlegel and R. Swinburne 155
5. S. Watanabe 157
Part II: Duration and Causality

VIII. The Intuitive Foundations of the Knowledge of Time 163


1. Preliminary Remarks 163
2. Intuitive Time and Memory 166
3. Intuitive Time and Perception 171
4. The Continuum of Instants Attached to Intuitive Time 174
5. Other Intuitively Knowable Aspects of Time 177
6. The Epistemological Significance of Intuitive Time 180
IX. Physical Time 187
1. General Remarks on the Causal Theory of Physical Time 187
2. Symmetrical Causality 197
3. The Concept of Event 205
T ABLE OF CONTENTS xi
4. The Ordinal Concept 209

4.1. Coincidence 209


4.2. The Spatio-Temporal Order 210
4.3. Simultaneity 211
4.4. Succession 213
5. The First Group of Axioms 214
6. The Geometrical Concepts 216
7. The Second Group of Axioms 217
8. The Third Group of Axioms 220
9. The Time Metric 223
10. Reference Systems, Coordinates, Transformation Formulas 226
11. Inertial Systems 230
12. Optical Criteria 234
X. Non-Physical Time 237
1. Psychophysical Time: The Perceptual Method 237
2. Psychophysical Time: The Expressive Method 240
3. Psychological and Inter-Psychological Time 248
4. Universal Time 251
5. Conclusion 259
Supplement

1. The Present Empirical Status of Psychophysical Parallelism 261


2.Conceptual Analysis of Psychophysical Parallelism 269
NOTES 286

INDEX OF NAMES 300


VOLUME II
TIME IN A QUANTIZED UNIVERSE

Part III: An Indeterministic Theory of Time

I Philosophical Interpretations of Quantum Physics


II The Problem of Causality in an Indeterministic Science
III Relativity and the Atom
IV Laws of Nature and Time's Arrow
V The Symmetry of Time and the Branch Hypothesis

Part IV: Universal Aspects of Time

I The Measurement of Time


II The Ontological Status of Time
III The Reality of Time
IV The Causal Nature of Time
V The Symmetry of Time
VI The Psychology of Time

Conclusion

Bibliography of Works Cited in Volumes One and Two, edited by Carolyn R. Fawcett

Bibliography of Writings of Henry Mehlberg

Index of Names to Volumes One and Two

xii
PREFACE

An intermittent but mentally quite disabling illness prevented Henry


Mehlberg from becoming recognized more widely as the formidable scholar
he was, when at his best. During World War II, he had lived in hiding under
the false identity of an egg farmer, when the Nazis occupied his native
Poland. After relatively short academic appointments at the University of
Toronto and at Princeton University, he taught at the University of Chicago
until reaching the age of normal retirement. But partly at the initiative of his
Chicago colleague Charles Morris, who had preceded him to a 'post-
retirement' profes-sorship at the University of Florida in Gainesville, and
with the support of Eugene Wigner, he then received an appointment at that
University, where he remained until his death in 1979.
In Chicago, he organized a discussion group of scholars from that area as
a kind of small scale model of the Vienna Circle, which met at his apart-
ment, where he lived with his first wife Janina, a mathematician. It was
during this Chicago period that the functional disturbances from his illness
were pronounced and not infrequent. The very unfortunate result was that
colleagues who had no prior knowledge of the caliber of his writings in
Polish and French or of his very considerable intellectual powers, had little
incentive to read his published work, which he had begun to write in English.
I vividly recall the sense of intellectual adventure I felt, when I first read
his beautiful long essay on the causal theory of time, while a graduate
student. It made me seek him out when he came to Princeton, and sub-
sequently to urge his appointment at Chicago. The essay had appeared in
French, in the Polish Studia Philosophica during 1935-37. It was a pity
that this remarkable piece of work has not been available in English until
now, when it is appearing thanks to Robert Cohen's unflagging initiative and
patient prodding of Mehlberg over a period of years.
This essay, and the succeeding contributions that are contained in the present
two volumes, bespeak Mehlberg's astonishing erudition and impressive
incisiveness: An extraordinary literacy in the history of philosophy, technical
physics, and modern logic, in the tradition of Polish logic at its best. Indeed,
Mehlberg'S command of the most recent relevant developments in theoretical
physics was outstanding even within the relatively small circle of philosophers
xiii
xiv PREFACE

working in the foundations of physics, most of whom are better known than he.

Hence it is my fond hope that the present two-volume opus will reach a much
wider public of scholars and students than his work has heretofore, to the
substantial enrichment of subsequent research in the field.

University of Pittsburgh ADOLF GRUNBAUM


EDITORIAL NOTE

Henry Mehlberg prepared the long introductory essay to these two books
which comprise his integrated life-long work on the nature of time and on the
philosophically pertinent and penetrating developments of theoretical physics
in the twentieth century. This essay, undertaken over some years, was to be
the last of his finished creative works, for he died in December 1979 while
these books were that the typesetter. He was contented, and delighted that the
fme work of his youth, the Essai of 1935, would come to be known again,in
his own revised and enlarged version; it is the principal text of the first
volume. The remainder of the two volumes consists of articles, some
previously unpublished, dealing with the relevance of quantum physics to the
problem of time. Written as they were in the decades since the Essai first
appeared, these articles include later views which at times are at variance
with some tenets of the earlier work; thus Professor Mehlberg came to a far
more critical view of the 'relational theory of time', and he devoted
considerable effort both to adapting causal theories of time to quantum
indeterminism, and to more deeply understanding the psychological aspects
of time, which he termed 'non-physical time'.

* * *
Mehlberg would have wanted to thank those who helped with this volume of
the Boston Studies in the Philosophy of Science: Paul Benacerraf for his
fme draft translation; Carolyn Fawcett for the Bibliography of Mehlberg's
writings and the Bibliography of works cited by Mehlberg throughout these
papers, and for her care in helping to complete the translation of the Essai;
Ellen Haring for her sympathetic but finn support through all the Florida
years; Mark Dinaburg for his splendid research assistance; Adolf Griinbaum
for his critical appreciation and warm-hearted encouragement; and above all
Susie Clark Mehlberg for her loving and devoted help, through happy times and
through troubled times. As friend who is also editor, I want to express my own
gratitude for such admirable and unselfish aid as this good man received in his
last decade while these books were underway.

* * *
Henry (Henryk) Mehlberg was born October 7th, 1904. A Polish Jew, edu-xv
xvi EDITORIAL NOTE

cated in Poland, a student of Ajdukiewicz, he was one of the great inter-war


generation of teachers and students in physics, logic, philosophy of science,
and exact philosophy. His parents were victims of the Holocaust, murdered in
Nazi concentration camps; his first wife, his beloved Josephine, herself a fine
mathematician, was a hero of the Polish underground forces. Mehlberg
survived the nearly five years of Nazi authority by moving to another city,
where he was able to obtain the identification card of 'Peter Justin
Suchodolski, agricultural engineer' from an influential friend. He also ob-
tained a suitable job which required him to absorb almost instant and fluent
knowledge of the composition of soils and fertilizers, as well as to register
with the SS Polizei; he had living quarters with an aristocratic Polish
landlady who preferred speaking French to Mehlberg and Josephine.
By his own account, Mehlberg had rejected the religious beliefs of his
mother and father by age 14. He said he became an atheist because the polish
language uses the same word for 'heaven' and for 'sky' (niebo):
' ... when I read the first sentence in the Book of Genesis, 'In the beginning
the Lord created the sky and the earth', I know positively that the sky is an
optical illusion and that the earth was not at the 'beginning' because it is
younger than the sun, so I then decided to study philosophy in order to get a
substitute for religion, and also Romance languages in order to make a
living.' (from a letter of December 21, 1976 to R.S.C.)
After his early post-war papers were published, Mehlberg emigrated to
Canada, becoming a Fellow of the Lady Davis Foundation, and teaching at the
University of Toronto. His appointment there was aided by a favorable review of
his Essai by a man he regretted never meeting, the American Julius Weinberg. In
his further emigration, his fmal one which was to the United States, Mehlberg
was at first stopped by the hesitation of an American consul in Toronto to allow
an immigrant from behind the 'Curtain', and especially during the pervasive
influence of the junior Senator from Wisconsin. It was the American philosopher
Sidney Hook who intervened successfully with the U.S. Department of State; and
it was again Professor Hook who persuaded Mehlberg to join the American
delegation to the 1952 Congress for Science and Freedom held at Hamburg.
Mehlberg's keynote address was a proud event for him; he was applauded,
particularly by the several dozen Nobel Prize winners, he felt, because of his
claim that any problem which could not possibly be solved by the scientific
method, has no solution. Bitterly criticized by religiously oriented philosophers,
and defended by such figures as his deeply respected former teacher Alfred Tarski
and by that noblest of German physiCists, Max von Laue, Mehlberg had risen to
a new
EDITORIAL NOTE xvii

level of public participation. His book, The Reach of Science, was the
development of his Hamburg lecture. And he was asked, along with the Spanish
philosopher Salvador de Madariaga, the Austrian physicist Hans Thirring, and
the German Social-Democratic Party leader Carlo Schmidt, to speak to the
German nation about the issues of the Congress over the German State radio
network. The others spoke in German; Mehlberg feeling still the potential victim
of the Holocaust, felt constrained to speak in English. In his last years, he
planned to write an autobiographical essay entitled 'Autobiography of a
Wandering Jew'. But he liked it when a friend told him that this was 'an American
success story', and that might have been the sub-title.

On reflection, Mehlberg considered his scientific obsession to be the problem


of time-reversal, his emotional puzzle the problem of a secular Jewish outlook in
an anti-Semitic atmosphere. Ever creative, but a troubled victim of his times,
Henry Mehlberg disciplined himself with astonishing success, for he overcame
political horrors and their derivative private maelstroms. These books are a
memorial to a splendid philosopher of science.

Center for Philosophy and History of ROBERT S. COHEN


Science, Boston University
September 1980
ACKNOWLEDGMENTS

The Essay which comprises Volume I of this work is based on a previously


unpublished translation, by Paul Benacerraff, of Essai sur la theorie
causale du temps which first appeared, in French, in Studia Philosophica 1,
pp. 119-
260 (1935) and 2, pp. 111-231 (1937), published by PAN (the Polish
Academy of Science), and reprinted here by their kind agreement with the
author.
In Volume II, Chapter II (,The Problem of Causality in an Indeterministic
Science') is reproduced from the International Journal of Theoretical Physics
2, pp. 351-372 (1969), and is reprinted here by kind permission of Plenum
Publishing Corporation. Chapter III (,Relativity and the Atom') first appeared
in Feyerabend, P. and Maxwell, G. (eds.) Mind, Matter and Method, pp. 449-
491 (1966), and appears here by kind permission of The University of
Minnesota Press. Chapter IV ('Laws of Nature and Time's Arrow') is adapted
from 'Philosophical Aspects of Physical Time', reprinted by permission of
The Monist, 53, pp. 340-384 (1969), and from 'Physical Laws and Time's
Arrow', from Feigl, H. and Maxwell, G. (eds.) Current Issues in the Philos-
ophy of Science, Copyright © 1961 by Holt, Rinehart and Winston. Chapter V
('The Symmetry of Time and the Branch Hypothesis') first appeared as
a review of the book The Direction of Time, by Hans Reichenbach, in The
Philosophical Review 71, pp. 99-104 (1962).
The Editor and the Publishers extend their grateful thanks to the
publishers of the above material for their permission to include it in the
present work.

xix
INTRODUCTION TO VOLUMES I AND 11*

There is no point in apologizing for the size of my work, nor even for the length
of the introductory chapter. A most enlightened thinker of the Age of
Enlightenment once stated that his volume would become much longer if he tried
to make it somewhat shorter. This illustrious comment also applies to my work,
although no 'enlightening' claims can be made for it. The tripartite title 'Time,
Causality, and the Quantum Theory' is meant to convey the order of priorities in
the ensuing discussions: the central topic is the problem of time. The issue of
causality will be explored in much detail in order to justify my principal
conclusion referring to the causal nature oftime, and it will also be imperative to
explore the many quantum theories, which originated in 1900 with Max Planck's
discovery of the quantum of 'action' (Le., the product of energy and time)
involved in any physical process. Planck generated a never-ending sequence of
quantal theories which, at a later time, have transcended the scope of physics
proper and invaded both chemistry and biology. The decisive relevance of this set
of theories to my causal approach to time is due to the fact that they both modify
the meaning of the Principle of Causality and shed new light on many aspects of
the issue of time, in addition to its causal nature.

In the context of the ensuing discussions, the temporal problem deals with
universal time, rather than with physical time only. It is obviously not the case
that only physical events occur in time. So do psychological events. Temporal
relations of simultaneity or succession (in addition to temporal patterns involving
events spread over temporal intervals of various lengths) may obtain among
exclusively physical events, or exclusively mental events (either intra-personal or
inter-personal) or among events classifiable under either category. One can
therefore state that, in addition to physical time, to intra-subjective and inter-
subjective psychological time, there is also a psycho-physical time. All these
varieties of time are components of universal time, and all are relevant in the
context of the ensuing discussions.
Accordingly, in addition to time-related fmdings of relativistic and/or quantal
physical theories, the psychological fmdings concerning time will
1
2 INTRODUCTION

also be discussed. They are not as conspicuous as the physical findings, and,
perhaps, not as revelatory. But they are by no means negligible and are actual-ly
necessary in any comprehensive account of universal time. Psychological
theories of time, to be discussed in several chapters of this work, deal with the
subjective (perceptual or memorative) awareness of things temporal, e.g., the
relations of simultaneity or succession, the length of temporal intervals and, in
addition, to the ideas of past, future and 'specious' present in the psycho-physical
area. These theories also contain findings concerning the relationship between
time-experiences in wakefulness and the dream-contents of sleep, in conjunction
with comparative findings of time experienced by normal and abnormal (e.g.,
schizophrenic or paranoid) adults. We shall also come to realize the significance
of neuro-physiological research about ex-perienced time, including the cerebral
localization of temporal experiences. Above all, the relevance of extra-physical
time to universal time will have to be explored.

At first, some comments are indispensable about languages and linguistic


formalisms inherent in physical or psychological sciences and their philosophy.
A preliminary meta-linguistic clarification of the issues involved in any science
and its philosophy is imperative, because linguistic confusions are known to be
conducive to inconsistencies and similar, dangerous outcomes. It is obvious that
even a talk about grade-school arithmetic is bound to be couched in a meta-
language different from the once Arabic, and now international, nota-tional
formalism of this elementary branch of number-theory. In this case, the meta-
language is spoken English.
It is a standard procedure now to classify the language used to formulate
fmdings about a scientific theory T couched in a language L as a meta-language
L' while L is called the object language of T. More specifically, if the meta-
language L' is associated with the object-language L, then, in order to enable L' to
serve in explorations of the intrinsic structure of L (i.e., its 'logical syntax') in
addition to investigations into the linguistic community using L ('the pragmatics
of L ') and the objects talked about by the users of L (the 'semantics of L '), then
the meta-language L' has to meet a set of requirements most of which are
satisfied on what A. Tarski would call the first semantic meta-language of L. L'
must qualify as the first semantic meta-language, because it often proves
necessary to study the tripartite properties of L' in the first semantic meta-
language of L', or the second semantic meta-language of L. The hierarchy of
semantic meta-languages associated with the object-language L may have to be
continued to consecutively higher ranks.
It is important to realize that a distinct hierarchy of meta-languages
INTRODUCTION 3

associated with an object-language L of a scientific theory T is often needed to


fully express T. At a later stage, we shall have to explore in detail a vital
dichotomy related to the classification of all meaningful statements of L into fact-
like and law-like statements. Roughly speaking, a statement S of L is law-like if
it specifies a universal regularity of entities dealt with in L, i.e., a regularity
which holds throughout cosmic space and time. The spatio-temporal universality
of law-like statements will also be shown to be associated with a conceptual
universality and to some extent, with quantificational universality. A true law-like
statement is usually called a natural law . Similarly, fact-like statements deal with
some local conditions prevailing in a fmite, spatio-temporal region. If true, a fact-
like statement is called a fact. Now, the important point is that, in addition to the
universal laws, formulated on the object-language L of a theory T, there often are
law-like statements belonging to the theory T but not expressible in L, because
they deal with the entire class of the law-like components of T which are
expressible in L. Thus, Einstein's Principle of Special Relativity claims that all
the laws of a relativistic physical theory T, which hold throughout the pseudo-
Euclidean, 4-dimensional space-time continuum, remain unchanged when a set
of spatio-temporal co-ordinates associated with some inertial frame of reference
is replaced with any other set of such coordinates obtained from the first set by a
transformation belonging to the Lorentz group. There is no noticeable advantage
in using the impressive term 'principle' instead of the humbler term 'law.' The
important point is that the Principle of Special Relativity deals with the aggregate
of all laws expressible in the object-language L of the theory T and therefore
requires a meta-language i associated with L. i is obviously not

a semantic meta-language of L, because the universe of discourse of i consists


of statements expressible in L. We shall call i the first, theoretical meta-
language of L and we shall not preclude the possibility of a hierarchy of
theoretical meta-languages of L. The distinction between the semantic and
the theoretical meta-languages of any object-language will be commented
upon in some detail at a later stage of my investigation. But both the
meaning of, and the need for, this distinction should be reasonably clear at
this pre-liminary stage of our inquiry.

2. PHILOSOPHY OF PHYSICAL SCIENCE vs. PHYSICS PROPER

Introducing a work dealing mostly, although not exclusively, with the phi-
losophy of physical science has little in common with the social event referred to
as the introduction of a human individual to another individual or group.
4 INTRODUCTION

At this juncture, it may be more helpful to quote a statement sometimes


attributed to Ernest Rutherford. This statement claims that there are only two
groups of sciences, viz. physics and stamp-collecting. If this view were really
attributable to the brilliant British scientist then its motivation would be
pretty obvious: Lord Rutherford, who came from the Far East to be fIred by
his incompetent colleagues at a Canadian university, spent most of his life in
the Cavendish Laboratory of Cambridge University in England and all the
momentous problems which he succeeded in solving there pertain clearly to
physical science. He may have under-estimated the import of Einstein's
theory of Special Relativity because it did not originate in Cambridge. I do
not want to detract from the unique status of the British city of Cambridge in
the history of science: the most spectacular episode in the entire history of
human science did happen in Cambridge, because the unique breakthrough in
both the history of physics and mathematics occurred there, owing to the
activity of one man: Sir Isaac Newton. But there is no point in comparing the
achievements of some incomparable people. All that matters at this introduc-
tory stage of our investigation is the alleged monopoly of physical science. It
would become obvious that imperialistic ideas about the range of physics are
untenable, if I could afford, at this juncture, a thorough analysis of the relevant
issue.
No such analysis will be attempted in this study, because of shortage of
space. But it seems sensible to state clearly, from the very outset, that the
investigations presented in the ensuing chapters of my work will deal with
the philosophy of physical science almost exclusively. The monopoly which
Lord Rutherford may have granted to his favored discipline would certainly
be unwarranted. But the strategic role which physical theories have come to
play in man's outlook justiftes the concentration of an investigation into
philosophical aspects of scientific time on the science of physics.
The present role of physical science involves both its relevance to
advanced technology and to several other fundamental sciences, including
chemistry, biology, astrophysics, and from a methodological point of view,
even econom-ics and literary history. The connection of physics and
advanced technology is so tight now that there would be no point in trying to
separate them. This implies that mankind, if deprived of advanced physics,
would be wiped out in a matter of days because it is impossible to feed four
billion human beings without the resources of an advanced technology,
inseparable from physics, as of now.
The relevance of physical science to man's general Qutlook is hardly con-
troversial. The beliefs we now hold about the universe we live in, its age,
INTRODUCTION 5

evolution, present composition and its pervasive, spatio-temporal continuum,


including the origin and nature of life, have originated with the surge of physics,
roughly simultaneous with the cultural surge of the Renaissance, about four
centuries ago. The importance of physics for the advance and understanding of
chemistry is best illustrated by the explanation of the chemical bond in quantum
theoretical terms, given by Linus Pauling. A more dramatic illustration of what
physics means to chemistry has been provided more recently by the discovery of
the spatial arrangement of atoms in a DNA molecule, along a double helix, by
Watson and his British associates. Perutz's ftndings concerning the hemoglobin
molecule are equally impressive.
In the vast area of biological sciences, which are likely to be gradually
absorbed by physics, the present relevance of physical science is conspicuous in
molecular biology, now sweeping the country. Obviously, the statistical methods
characteristic of most of physical research have been dominating a crucial area
of biology, viz. genetics, decades before the fad of molecular biology came into
being. Similar statistical methods have also invaded a large territory of
economics where they go by the name of 'econometrics.' John von Neumann, one
of the greatest mathematici~s of this century, did a pioneering job in this area too.
The applicability of typically statistical methods to literary history is illustrated
by the problem concerning the chronology of Plato's dialogues. The issue has
been controversial for centuries. It seems reasonaly settled at present, due to the
application of statistical (,stylometric,' in this case) methods of Campbell and
Lutoslawski.
It is not easy to draw a reasonably clear line separating physics proper from
the philosophy of physics. The difftculty is partly due to the circumstance that
important contributions to the philosophy of physics have often been made by
leading physicists. This was the case of Ampere and Poincare in France, of
Bridgman and Wigner in the United States, of Lord Kelvin and A. S. Eddington
in England, of Niels Bohr in Denmark, etc. The point is, however, that these
physicists (and astronomers), who have substantially enriched the philosophy of
physical science, seem to have always been clearly aware of the difference
between physics proper and its philosophy. Moreover, signiftcant advances in the
philosophy of physics have frequently been due to philoso-phers who have never
been practitioners of physics. Such was the case of Carnap and Reichenbach in
the United States, Bergson and Meyerson in France, Popper and Russell in
England.
To determine the intrinsic difference between physics proper and its
philosophy, the above details concerning the division of labor in both areas are of
little interest. A survey ofthe basic and typical problems in the philosophy
6 INTRODUCTION

of physics, in contrast to problems dealt with in physical science, would


certainly be more helpful. At this juncture, we shall confine ourselves to
those strictly physical problems which pertain to a strategic area of physical
research, viz. the aggregate of currently available quantum theories. This
aggregate includes Bohr's obsolete version of quantum theory, followed by
the theory of non-relati'listic quantum mechanics originated mainly with de
Broglie, Heisenberg and Schrbdinger (1924-1926). To provide a
representative sample, this theory must be supplemented by relativistic
quantum electro-dynamics whose present shape is due mainly to Dyson,
Feynman, Schwinger and Tomonaga. In addition, we have to list the
quantum theory of elementary particles, overlapping with relativistic
quantum field theory. In this case, the representative scienti3ts are legion.
Some may nevertheless be quoted: Gell-Mann, Wigner, Yukawa. However,
Wigner's contributions have appreciably affected the entire area of quantum
physics, as did Dirac's contributions. Quantum statistics, either of the Bose-
Einstein, or the Dirac-Fermi type, should be added to this small sample.
The characteristic problems dealt with in the above as well as in related
quantum theories can be described as follows: the entities referred to in these
problems have several, basic properties of the 'quantized' variety. This means
that the physical objects or systems (Le., either elementary particles, or
nuclear, atomic, molecular and other systems of elementary particles)
described in the relevant quantum theories possess basic quantities which are
incapable of a continuous change but have to be treated as N-tuples of
appropriate units. Typical examples of such quantized properties are
provided by the electric charge, the leptonic or baryonic charge, the
'strangeness,' the 'spin,' etc. Even 'charm' seems to be such a property of the
sub-particles of elementary particles, called 'quarks', in Pasadena, California,
and currently as problematic as any charming entity could possibly be. The
spatio-temporal arena of the cosmic drama is not quantized so far.
The types of entities dealt with in contemporary quantum theories clearly
suggest that quantum physics is an empirical discipline. This does not preclude
an appreciable use of deductive methods but the ultimate body of evidence in
support of either a single (law-like or fact-like) statement, or of a complex
theory, is always observational. If the supporting observational evidence is
obtained by using a special arrangement of physical objects produced for this
purpose (Le., with a view to either solving a pre-assigned, physical problem, or
to checking on a proposed solution to such a problem) then the observational
evidence is usually promoted to the coveted rank of 'experimental evidence'. In
both cases, the indispensable role of observational evidence is evident. We
INTRODUCTION 7

have thus obtained a very sketchy account of both the subject-matter and the
method ascribable to quantum physics proper.
The philosophy of physics differs from physics proper in these two crucial
respects. The subject-matter of this branch of the philosophy of science has
nothing in common with that of physics. The philosophy of physics deals
with a class of entities foreign to the physical universe, viz., with the set of
presently available physical theories. In other words, the subject-matter of
the philosophy of physics is entirely man made, whereas, in all physical
theories, man is a tiny item of the biosphere, which, in turn, isjust an
appendix of the colossal range of physics, both experimental and theoretical.
Accordingly, a typical problem in the philosophy of physics does not look for
observational support. To give an idea of methods adequate for the
philosophy of physics, I would like to emphasize the importance of an
extremely felicitous achieve-ment of contemporary physical science.
I have in mind the fact that all major physical theories, including all the
quantum theories of interest to us, have been finitely axiomatized. To explain
the relevance of this achievement to the philosophy of physics, let me first
point out that every theory of this type can be construed as a deductive
system, this means as an infmite set of statements all of which are logical
consequences of a subset of this set. A statement s is said to be a logical
consequence of the set S of statements if, on supplementing S by the two basic
logical calculi (the propositional and quantificational calculi), we can obtain s by
applying to the expanded set S a finite number of times the 'rules of inference'
valid in the logical formalism of the physical theory under consideration (e.g.,
the rule of substituting appropriate constants or variables in a formula already
derived in this way from the extended set S). (In math-ematical parlance, a
deductive system is a set of statements, closed under arbitrarily long applications
of the logical rules of inference.) Accordingly, we must not confuse a deductive
system of statements with an axiomatic system of statements, even if the same
statements occur in both systems.
An axiomatic system representing any pre-assigned, deductive system D, can
usually be obtained by selecting, in the deductive system D: (1) a finite set of
terms (primitive terms) to be used in the axiomatic system although no definition
is made available for them. (2) a fmite set of state-ments (axioms or postulates)
to be used as premises in the proofs of any other statement of the axiomatic
system. A deductive system D in which such two selections can be effected is
said to be finitely axiomatizable, provided that two more fmitude requirements
are met in D; (3) all proofs of non-axiomatic statements of D are of finite length.
(4) all meaningful statements
8 INTRODUCTION

in D are finite sequences of terms definable by means of the primitive terms


available in D.
None of these four requirements of fInitude is tautologically satisfIed.
Interesting results have been obtained in metamathematical investigations by
dropping some or all the finitude requirements. For instance, GOdel's
Incompleteness Theorem can be dodged in this way. However, the signifIcant
fact concerning the aggregate of all contemporary physical theories (Le., the
subject-matter of most of my ensuing investigations) is theirfmite
axiomatiza-bility. In most cases I am referring to an actual aspect of these
theories, and not to their potential axiomatizability. Thus, Newtonian
mechanics was presented by Newton himself in an axiomatic form, viz., as
the set of his three Laws of Motion. A more modern ('formalized') axiomatic
system for Newtonian mechanics has been worked out by P. Suppes. In the
time-interval separating Newton's work from the fmdings of a contemporary
logician, a different, most effective and promising axiomatization of classical
mechanics was obtained by Maupertuis, who succeeded in squeezing the
postulational basis of Newton's theory into a simple, concise principle, viz.
the variational Principle of Least Action. The teleological interpretation put
by Maupertuis on his variational Principle did not survive. But the variational
principle came to be embedded into the calculus of variations, which in turn
was included in the comprehensive theory of functional analysis, an
indispensable, mathe-matical tool of contemporary theoretical physics.
The availability of a concise axiomatization holds also for Maxwell's
theory of the electromagnetic fIeld with a non-vanishing density of electric
charge and current. Phenomenological thermodynamics has been superbly
axiomatized by C. Caratheodory, who also worked out a set of axioms for
Einstein's Special Relativity. Axiomatic systems for Special and General
Relativity are due to H. Reichenbach. The axiomatization of non-relativistic
quantum mechanics is due to J. von Neumann, and,in a more modern
fashion, to' G. Mackey. Relativistic quantum mechanics of electrons and
positrons has been presented, in an axiomatic form, by P. A. M. Dirac. An
elegant axiomatization of quantum electrodynamics in creditable to R. P.
Feynman. Relativistic quantum fIeld theory was fIrst axiomatized by
Lehmann, Szy-manzyk and Zimmermann, and, subsequently, by A. S.
Wightman. Haag and Ruelle succeeded in axiomatizing the quantum theory
of ('asymptotic') elementary particles.
The axiomatization of signifIcant, physical theories can be considered as
one of the principal tools of the philosophy of physics. However, in contrast
to axiomatic systems of purely mathematical theories, the axiomatization of
INTRODUCTION 9

a physical theory also requires a supplementation of the axiomatic formalism


with a set of 'defmitional criteria', or 'operational definitions'. These defmi-
tional criteria establish a link between the mathematical formalism of the
theory and the observational results which may substantiate or disconfirm
the theory. In contrast to classical definitions of the equivalence type which
provide an expression already available in the language and interchangeable
with the defmed term, the definitional criteria and the operational definitions
contain only an expression already available which states a sufficient, but not
a necessary condition for the applicability of the defined term, and is,
consequently, not interchangeable with the latter. Thus, 'a rectangle with
equal sides' defmes a 'square' in the classical way. Both expressions are inter-
changeable. 'Deflecting a magnetic needle' is merely a definitional criterion
of an 'electric current'. The two expressions are not interchangeable because,
in the absence of a magnetic needle, an electric current may still flow.
I have shown elsewhere that this crucial definitional technique of physical
science is largely responsible for its ability to transcend the scope of pre-
scientific knowledge and even to provide for new, irreducible information in
every physical theory which derives its defmitional criteria from some previous
physical theories. Thus, thermodynamics is not reducible to mechanics,
electrodynamics transcends the former two theories combined, special and
general theories of relativity go beyond the aggregate of pre-relativistic theories,
and new, cognitive quantum jumps occur from pre-quantal physics to the current,
quantal theories which have kept originating ever since Max Planck started the
twentieth century with his quantum of action.
Albert Einstein may have been the first to realize the role of definitional
criteria in those axiomatic systems which represent physical theories. In his
1905 pioneering paper on 'The Electrodynamics of Moving Bodies', this
insight is contained in his analysis of the concept of simultaneity. In 1927,
Percy W. Bridgman used Einstein's insight to work out a systematic
presenta-tion of defmitional criteria or operational definition. Needless to
say, only Einstein's genius could have enabled him to derive, from his
analysis of the meaning of simultaneity, the Special Theory of Relativity,
possibly the single most significant scientific advance of this century.
The finite axiomatizability of all major, current physical theories, calls for
one more explanation: what class of entities, what sub-universe of the physical
universe, is made cognitively accessible to man, by this aggregate of theories? In
the sequel, this question will be explored in detail. At this juncture it may suffice
to introduce A. de Morgan's terminology and to call the 'universe of discourse of
a theory T' the class of all individuals over which the theory
10 INTRODUCTION

makes any non-tautological statements. The constant symbols referring to any


single individual, and the variables which range over their totality are called the
individual symbols of T. To avoid self-contradictory statements in T, we have,
according to B. Russell's discovery of 'logical types', adequately to stratify the
non-individual constants and variables of T. The simplest way of constructing
such a consistent formalism was discovered independently by Godel and Tarski,
who had been in a position to take advantage of an earlier result of Wiener. Apart
from individual symbols and the standard sentential connectives, the Godel-
Tarski formalism contains only an existential or universal quantifier, in addition
to a class of one-place predicates of every finite, integral level N, with the
understanding that a predicate of level N denotes a class whose elements are of
level N-l and that predicates of level 1 denote classes whose elements are
individuals.
Apart from axiomatizing significant, physical theories, the philosophy of
physical science would have to determine the definitional criteria associated with
the theory and its universe of discourse. The other objectives of a phi-losophy of
physics consist in determining how the physical theory under con-sideration
affects philosophy's major issues, e.g., empiricism vs. rationalism, determinism
vs. indeterminism, epistemological or ontological idealism (sometimes called
'instrumentalism') vs. realism, materialism vs. spiritualism. Needless to say,
classical problems of ontology and epistemology take on a specific form when
explored in the philosophy of physics. Thus, the epis-temological problem of
determining the nature and scope of human knowl-edge, on being applied to the
quantitative statements of physics, gives rise to the essential problem of
measurement, particularly in quantum physics. In addition to the relevant work of
von Neumann and Wigner, the treatment of the problem of measurement by the
Yale University philosopher and physicist, H. Margenau, has achieved
international recognition. This holds also of A. Griinbaum's work on the
epistemology and ontology of space and time and of O. Costa de Beauregard's
work on time. At this juncture I shall merely mention the contribution to the
philosophy of physics made by Destouches-Fevrier and d'Espagnat in France, and
the achievements of M. Bunge, P. K, Feyerabend, E. Nagel and D. Shapere in the
United States.
The answer which I have suggested to the most natural question 'what is
physics about?' is obviously conservative. My preliminary answer states that
physics deals with matter, i.e., the sum-total of elementary particles and of
systems of elementary particles which are scattered throughout space-time and
either undergo metamorphoses transforming particles of one species into
particles of another one, or travel through space as time goes on. However,
INTRODUCTION 11

the answer is merely preliminary and must not be 'conservative' at all. The
radicalism of the answer comes mainly from the fact that some
fundamentally important particles have no intrinsic ('rest') mass at all, e.g.,
the photons and the neutrinos. In addition, some particles are not localizable
at all, i.e., that strictly speaking, they are nowhere. The photons fall under
this category. To what extent physical entities of this sort can be considered
as constituents of 'matter' depends upon the feeling of linguistic propriety. Sir
Isaac Newton would certainly turn down any attempt to ascribe the attribute
'material' to photons and neutrinos.
There is another, more serious objection to identifying the subject-matter
of physical research with matter even if mass-free matter were considered
unobjectionable. At a later stage of this investigation, we shall be dealing in
some detail with 'relativistic theories involving second quantization'. A
physical theory is termed 'relatiVistic' or 'covariant' ifthe axioms on which it
rests involve spatio-temporal coordinates and remain unchanged when these
coordinates are transformed under the Lorentz group. In the pre-quantal era
of physics the mere fact that the space-time coordinates of a theory
transformed under the Lorentz group was sufficient to guarantee that its
axiomatic laws remain then invariant. This is no longer the case in relativistic
quantum theories. In the latter case, of decisive importance in this investiga-
tion, the additional conditions required for the relativistic covariance of the
quantum theory under consideration have been determined and found true to
facts by the Russian mathematician Gel'fand. Thus, in the quantum case it
takes more for a theory to be relativistic than to have its space-time
coordinates transform under the Lorentz group.
The 'second quantization' of a physical theory T comes roughly to the
following: the physical condition prevailing in a spatio-temporal region R is
described by the 'quantum state' Q of R, i.e., (in the 'SchrOdinger-representa-
tion') a complex-valued function of the spatio-temporal co-ordinates of R.
The measurable quantities defmed over R are no longer represented by some
real-valued functions defined over R (e.g., scalars, vectors or tensors) but
rather by operators associated with a suitable Hilbert space and also mapping
Q into some other quantum-state Q. Usually, those operators are of the
'Hermitean' variety which, in some important cases, become occupation-
number operators. The particles located within R are then represented by the
'eigen-values' of the appropriate occupation-number operators. Thus, fmite
spatio-temporal regions have now the logical rank of individuals, ( = 1), the
rank of the quantum states is 2, the operators over the quantum states have
rank 3, and the particles are promoted to rank 4. Intrinsic properties of
12 INTRODUCTION

particles would be mapped into classes of particles and thereby acquire rank
5. Accordingly, the subject-matter of physical science has a stratified
structure, based on the spatio-temporal universe of discourse and rising to
ontological layers of increasing logical rank. This stratified hierarchy will be
explored in detail at a later stage of our investigation.

3. SURVEY OF PROBLEMS DEALT WITH IN VOLUME I

In the remaining sections of this opening chapter I would like to give the
reader an idea of how this work is organized and what main topics in the
philosophy of physical science are dealt with. [The work, in two volumes,
consists of four parts, Volume I containing Parts I and II, volume II contain-
ing Parts III and IV - Ed.] The principal objective of Parts I and II of Volume
I is the exploration of the philosophically significant aspects of the timeless
problem of time. This problem is geared to the puzzling temporal entity
without being confmed to it. This objective can be circumscribed somewhat
more precisely by stating, at the very outset, that only scientific uses of
temporal concepts will be explored. More specifically, I shall concentrate the
ensuing discussions on the physical and the psychological aspects of time.
The tempting biochronological issues and those concerning linguistic,
geophysical and cosmological time will be studiously avoided in most of my
work. The justification of this restriction rests on the fact that, on closer
analysis, the bulk of presently available, reliable and socially relevant
information about time is provided by physics and psychology.
Simultaneity, succession and length of duration are obviously central to the
problem of time. Accordingly, much of the sequel will deal with these
fundamental temporal relations. However, the so-called relational· theory of
time, which reduces temporal problems to those dealing with temporal relations
and which is outlined in the Essay on the Causal Theory of Time (Volume I
of this work), proved inadequate when an analysis of quantum the-oretical time
was called for. Accordingly, time will be investigated in Parts III and IV (Volume
II of this work) as an entity in its own right and will be found to provide the gist
of the reality of the physical universe we live in and the psychological world
somehow embedded in this universe. Einstein's relativistic time theory, espoused
in this work, is different from a relational theory and actually incompatible with
it. However, the investigation will deal with the philosophy of time rather than
with the physics and psychology of time.
The philosopher used to be mankind's spokesman whenever major scientific
advances materialized. This was still the case when Einstein's two relativity
INTRODUCTION 13

theories came into being. They were conunented upon by B. A. W. Russell


and A. N. Whitehead in Great Britain, P. W. Bridgman and H. Margenau in
the United States, H. Bergson and E. Meyerson in France, R. Carnap and H.
Reichenbach in Germany, M. Schlick in Austria and this writer in pre-war
Poland. However, this age-long social function of philosophy is no longer
discharged at present. This fact is a symptom of the cultural malaise we are
now experiencing and may be partly due to the reluctance of the world's
philosophical community to obtain a reasonable conunand of the consider-
able amount of indispensable technicalities involved in the strategic center of
contemporary science, i.e., the aggregate of current quantum theories
(primarily quantum physics, quantum chemistry and quantum biology).
These theories are, however, of vital importance for both man's general
outlook and his material survival. The atom and the hydrogen bomb, whether
used in peace or in war, were produced on the basis of quantum theoretical
information, by E. Fermi and E. Teller, respectively. The impetus was due to a
letter written by Einstein to F. D. Roosevelt and taken to him by Szilard. Fermi
made important contributions to the aggregate of quantum theories, particularly
to quantum statistics. All quanta, or elementary particles, are pre-sently classified
under the headings of fermions or bosons. Bose and Einstein were responsible
for an alternative quantum statistics. There are few features of the contemporary
scientific landscape that are not partly traceable to Einstein.
It is the purpose of the present work to make up for this philosophical
deficiency, at least to an infmitesimal degree. While Volume I of this work
contains a modified English translation of my Essai sur la theorie causale du
temps, I (1935), II (1937), Volume II, parts III and IV are new and were written
in English from the outset (between 1966 and 1969). In Part I, a chapter (Chapter
7) on more recent alternative approaches to the problem of the direction of time
has been added. The irmovation of the English Part II is mainly due to a
presentation and evaluation of the work of J. Eccles, D. Ingle, N. Kleitman, H.
Kliiver, W. Penfield and A. Rechtschaffen as related to the problem of time via
the mind-body issue. With regard to the specific solution to the mind-body
question usually referred to as 'Psychophysical Parallelism', I also attempt to do
justice to J. von Neumann's challenging reference in connection with his own
fmding concerning the arbitrariness of the 'cut' separating the measured object
from the measuring instrument. Similarly, the contribution of two biochemists
(von Foerster and Hoagland) to the cerebral localization of our intuitive
knowledge of time is taken into consideration. (These new sections of the
English Part II are included as a supplement following Chapter X, the last
chapter of Volume I.)
14 INTRODUCTION

Chapter VIII (in Part II of Volume I) deals with the intuitive foundations
of our knowledge of time, both physical and extraphysical. Formerly, I
construed these intuitive foundations mainly in terms of our ability to tell
perceptions from recollections and to perceive temporally extended patterns.
New insights into intuitive time are creditable to the psychoanalytic move-
ment. The time of schizophrenics, no less important than the time of normal
adults, was successfully investigated. The temporal threshold of auditory and
visual sensations was explored from new points of view by researchers who
were in a position to utilize the classical fmdings of Stevens and Waldo New
fmdings concerning the perception of temporal patterns are now available .
.The very nature of perception, closely related to, and yet different from
hallucination, is now better understood. If H. Taine were still alive, he would
no longer state that perceptions are true hallucinations. These hints may
suffice to give an idea of how the English Parts I and II differ from their
French counterparts.

4. SURVEY OF PARTS III AND IV (VOLUME II)

I shall now start listing the main topics of the first part of Volume II (i.e.,
Part III). The crucial, epistemological issue in the philosophy of quantal
science, is obviously the problem of measurement. This has already been
outlined in § I of this Introduction, in connection with the measurement of
time. In the fIrst four chapters of Part III, this issue is explored in more
detail, and an attempt is made both to do justice to the relevant, extant
literature and to propose an approach of my own, hopefully not as obsolete
as some of the extant literature already is.
Part III is essentially an indeterministic theory of time. In it, the discoveries
and insights of the set of contemporary quantum theories are closely studied. The
fIrst result is the surrender of strict determinism, incompatible with the central
quantum theories now available, e.g., the Heisenberg-Schrodinger non-
relativistic quantum mechanics, Dirac's relativistic quantum mechanics or the
Feynrnan-Schwinger quantum electrodynamics. The last theory is both
relativistic (i.e., invariant under a Lorentz transformation of the spatio-temporal
coordinates) and twice quant.ized (Le., implying the reinterpretation as self-
adjoined operators of the scalar- vector-, and spinor-valued functions of the
above coordinates). The aforementioned unavoidable technicalities consist in the
use of self-adjoint operators over Hilbert spaces, of several other chapters of
'functional analysis' and of the theory of continuous group re-presentations. The
men responsible for these innovations of the mathematical
INTRODUCTION 15

formalism of quantal theories are, respectively, John von Neumann and Eugene
P. Wigner.
In evaluating the single, philosophically most significant feature of the
quantal theoretical outlook - the surrender of strict determinism - both ontology
and epistemology are involved. However, the epistemological aspect, viz. the
restriction on predictability is secondary. The ontological impact is essential, not
only because of its relevance to one of philosophy's most vital issues, the
freedom of the will (which could not be discussed in my work). The scientific
world view is affected and the obvious task of a philosophical interpreter of
physical science is to limit the scope of the ontological disaster associated with
the breakdown of determinism regardless of whether or not it is relevant to the
freedom of human will. The exploration of the issue in both relativistic and non-
relativistic quantal theories yields the following conclusion: Although the claim
of a universal and strict determinism has to be abandoned at present, an
indeterministic extension of the idea of causality leads to a valid Indeterministic
Principle of Causality. The new Principle can render almost every service which
was credited to its deterministic forerunner. More specifically the causal
approach to the problem of time need not be abandoned because of the surrender
of strict determinism.
Actually, the breakdown of strict determinism in a quantized universe ls not
pervasive, because in a Significant class of physical processes, the fmal stage is
accurately predictable if the initial stage is known and the relevant quantal
system is adequately isolated (i.e., if any external interference with the system
during the unfolding of the process is negligible). Thus, in non-relativistic
quantum mechanics, the transition of an isolated physical system from its initial
to its fmal quantum state is governed by Schrodinger's partial differential
equation. The equation is of the first degree with regard to time and, accordingly,
establishes a rigorous, deterministic connection between the consecutive states of
the system. This holds also for the overall values of quantities governed by
conservation-laws and ascribable to an adequately closed system. Thus, the
conservation of energy and of linear momentum provides for an accurate
predictability of future, relevant values on the basis of earlier values. So far, I
have been unable to fmd out whether the range of strict determinism in the
quantal area would suffice for the substantiation of a causal theory of time. If it
did, then the Indeterministic Causality Principle would yield a welcome,
alternative approach. If it did not, then we would have to rely on this
Indeterministic Principle entirely.
In discussing the philosophical aspects of quantum time in Part III, the fust
fundamental issue is the ontological (or referential) scope of quantal
16 INTRODUCTION

time. An attempt is made to avoid the 'genetic fallacy' in determining the


reach of quantal laws involving time in an essential way. Granted, quantwn
physics got started at the micro-level. But, it kept expanding and the discov-
eries made during the many decades of 'the rise and fall' of quantal theories
leave no doubt about the universality of quantwn laws governing micro-
objects, macro-objects, astronomical and cosmological objects or systems
and super-systems of physical objects. They all involve the same temporal
component of Minkowski's four-dimensional world. This follows from the
availability of well-established quantum theoretical fmdings involving time
and applicable either to one or to several, or to all size-levels of the universe.
In the context of my investigations, the single, possibly most significant
fmding in the quantal area is the TCP Theorem, dealing with the sequence of
time-reversal, charge conjugation, reversal of spatial directions (parity). The
objection to the relevance of some processes dealt with in elementary particle
physics (viz. the fact that their alleged irreversibility is not law-like) does not
apply to the TCP Theorem. Since it is creditable to the insight of Schwinger,
. and has been established in several ways by Liiders and Pauli, there is no
point in questioning its present status, although some possible violations of
this invariance-principle have also been discussed. My main conclusion is
that time-reversal has to be explored in conjunction with space-reversal and
charge conjugation. This is just another case of the rule which we already
have dis-cussed in connection with the time-symmetry of all major, physical
theories: In Newtonian mechanics, it did not suffice to substitute -t for t. The
available, time-symmetrical nature of forces governed by Newtonian law had
also to be considered. Analogous, additional modifications, required to show the
time-reversal invariance of other physical theories, have also been mentioned.
Hence, the TCP Theorem fits into the overall temporal perspective inherent in
current, physical theories.
Apart from replacing strict determinism with an indeterministic version of
the principle of causality (in Chapter II, Part III: 'The Problem of Causality in
an Indeterministic Science'), several other aspects of the time problem are
dealt with in the consecutive chapters of this Part. They include, for example,
the replacement of Einstein's original, macrophysical idea of an inertial
frame of reference with a new concept valid within and without quantwn
theory. Later, the specific problems raised by quantum time are attacked.
These problems consist in the new status of time-symmetry and time-reality.
It would be unreasonable to expect that the representatives of the most
rational, empirical science (Le., theoretical physics, sometimes called 'applied
mathematics', or even 'natural philosophy', by users of The Queen's English)
INTRODUCTION 17

would refer to the problem of measurement in the same way. They do not. The
German names of the authors of a French monograph on the (interna-tional)
problem of measurement prefer to refer to it as 'The Theory of Ob-servation in
Quantum Mechanics'. Pauli and later, von Weizsacker introduced a terminology
of their own: 'The Problem of Objectifiability in Quantum Mechanics'. I shall
stick with the 'objectifiability' terminology, faute de mieux. The new twist in
this vital, mensural problem is probably derivable from Heisenberg's Uncertainty
Principle, which banished simultaneous sharp measurements of many pairs of
vital, 'conjugated' quantities, e.g., the vectors position and/or linear momentum
of a small particle at the same time. But an entire new philosophy, co-authored
by N. Bohr and W. Heisenberg, has somehow emerged in connection with the
Uncertainty Principle. To me, this philosophy seems outrageous and I am not
surprised at all that the name of this philosophy is geographical, rather than
human: 'the Copenhagen Interpretation'. The Danish city has an illustrious
history, human and not imperialistic. But the proponents of the universal validity
of the Copenhagen Interpretation represent an imperialistic policy in the
scientific repUblic. Needless to say, some of their followers, e.g., L. Rosenfeld,
went even farther.
Since we are interested in the philosophical implications of the main theories
of quantal measurements, we do not have to worry seriously about the physical
paradoxes which have probably kept professional physicists worried. Thus, the
impossibility of sharp, simultaneous measurements of conjugated quantities, e.g.,
position and linear momentum or time and energy has often been construed as
the ontological (rather than epistemological) claim that a particle whose position
has been 'sharply' measured, has no momentum: It is neither at rest, nor in a state
of motion in any direction and with any velocity. Conversely, if the momentum of
a particle has been sharply measured, then it has no position, it is nowhere. This
is the case of photons, whose velocity has worried scientists ever since
Michelson measured it, until Einstein's Special Relativity provided some relief
for them. From a philosophical point of view, particles dealt with in quantal
theories are not individuals, since they have climbed to the fourth logical rank.
But it is defmitely of philosophical interest, whether particles like photons exist
at all. Denying this would be ontological folly, since photons can be counted.
Denying our ability to know anything about photons would be an epistemo-
logical disaster. Several scientists have tried to remedy the disaster of the
Copenhagen Interpretation by liberalizing the relevant views proposed in
Copenhagen. From a critical survey, I have concluded that they did go in the
right direction but did not go far enough. Eventually, I shall try to spell out
18 INTRODUCTION

the principal epistemological and ontological implications of an acceptable


Quantum Measurement.
These two-fold implications will be termed, respectively, empiricist epis-
temological and empiricist ontological realism. The position of empiricist
epistemological realism in the philosophy of a set of scientific theories
comes to the claim that these theories provide scientific knowledge (Le.,
socially relevant information) adequately supported by available,
observational and publicly verifiable evidence about observer-independent,
though observable, entities dealt with in these theories. The set of observer-
independent entities referred to in the aforementioned theories includes the
class of all finite, spatio-temporal regions, the quantum states prevailing in
these regions, the mappings of these quantum states into other quantum states
effected by asso-ciating these states with some relevant Hilbert spaces and
defining appropriate operators in these spaces, mainly the Herrnitean variety.
In addition to these three consecutive layers of physical reality, empiricist
epistemological realism also claims the availability of scientific information
about observer-independent particles, represented by the eigen-values of the
relevant occu-pation-number operators in addition to several physical
properties like electric charge, rest-mass and spin ascribable to these
particles. For the sake of consistency and freedom from antinomies, five
consecutive logical ranks have to be assigned, respectively, to the class of
spatio-temporal regions, followed by the class of quantum states, which in
turn leads to the class of operators; then the class of eigen-values of
occupation-number operators and fmally the cognitively accessible
properties of the particles corresponding to these eigen-values.
We thus reach five levels of knowable physical entities. Only the first level
consists of individuals, the other four levels are made up by classes of various
logical ranks. The epistemological claim that all these strata are within the reach
of scientific knowledge may be classified as an empiricist epistemological anti-
nominalistic realism. Similarly, the claim that all these five types of entities are
observer-independent and exist falls under the category of an empiricist,
ontological and anti-nominalistic realism. Anti-nominalists of any denomination
are often referred to as varieties of Platonists. However, the Platonic label is very
misleading, because of the connotation of idea entities, or essences, usually
ascribed to Platonism. This is why I prefer to label my position as an empiricist
anti-nominalistic realism (either of the epistemo-logical or the ontological
variety). Nominalism used to be fashionable a few decades ago, perhaps because
it was jointly advocated for a while by the impressive enterprise of Goodman and
Quine. However, Quine came to realize
INTRODUCTION 19

the untenability of the nominalistic position even in the logico-mathematical


area. And Goodman then claimed that the anti-nominalist outlook, which he still
labelled as Platonism, is "poison." It may be that the advocacy of an anti-
nominalist, empiricist position is somewhat risky, but I shall try to show that this
version of anti-nominalism is nevertheless valid and justifiable. Those who are
afraid of the alleged poisonous effects of an empiricist anti-nominalist view will
have to realize that it may be somewhat risky to profess the truth in many areas.
There is no risk-proof approach to truth, neither in the prosecu-tion of criminals
nor in interpreting current, scientific theories.
The complex issue of an anti-nominalist realism has been, still is and,
probably will be discussed over and over again by thinkers who are interested in
these areas of epistemology or ontology. I shall not try to explore these issues
outside of the epistemology and ontology of phYSical science and, more
specifically, of the set of currently available quantal theories. My only objec-tive
will be the presentation and evaluation of the new arguments which can
allegedly be derived from these theories in support of an anti-realist position.
And my conclusion will claim the fallaciousness of all these arguments.
In any event, we have to realize that the ostensibly non-philosophical issue of
the nature of quantal time-measurement may have significant implications of
both an epistemological and ontological nature. This holds also of the many
alternative approaches to quantal measurements which are not geared to time.
The main objective of the first four chapters of Part III is to deter-mine these
philosophical implications. Obviously, this could be achieved without exploring
the purely scientific issue sufficiently and, consequently, without presenting
several technicalities, or preferably, several intrinsic features of quantal
measurement. I do not think that the price (of plunging into some pretty technical
discussions) is exorbitant, if the plunge should yield some philosophical insights.
After all, the philosophy of science is rational talk about science, with a view to
getting answers to philosophical queries. If we do have to talk about science then
we would certainly prefer to know what we are talking about.

Then, treatment of the interpretational problem in the philosophy of quantum


physics is presented. The philosophical interpretation of any em-pirical theory is
bound to deal with both the cognitive foundations of the theory and its
ontological scope, i.e., the class of entities the knowledge of which is provided to
the theory. In traditional parlance, both the epistemo-logical and the ontological
relevance of the theory have to be determined in a valid interpretation.

I shall try to show that the many constructive suggestions made in the
20 INTRODUCTION

relevant literature contribute to restrict the anti-realist claims inherent in the


Copenhagen Interpretation, but fail to deal adequately with all these claims.
There is no point in dealing, within the context of my work, with the entire
problem of realism versus idealism. I shall simply examine the principal new
arguments which are frequently derived from quantum physics in support of the
Copenhagen Interpretation. The conclusion derivable from a critical analysis of
these arguments will be as follows: all these arguments, examined jointly, do not
weaken the position of an epistemological and ontological realism, in its
empiricist and anti-nominalist version.
In the main, the quantum theoretical arguments which, ostensibly, tend to
weaken the aforementioned alliance of epistemological and ontological
strongholds, are related to the following, scientific fmdings:
(1) The duality of particles and waves, started by the Einsteinian, corpus-
cular theory of photons, which apparently was compatible with the undulatory
theory of electro-magnetic radiations traceable to Fresnel and Maxwell, was
extended by L. de Broglie and became a pervasive feature of the physical
universe. The standard, anti-realist argument, based on the particle-wave duality,
claims that the entities described and explained in quantal theories, could not
possibly exist, because they would have to be both corpuscular and undulatory,
which is impOSSible. Thus, in most cases, a particle is sharply localizable while
the corresponding wave would unfold itself in a spatial volume whose order of
magnitude is incompatible with any sharp localization and may coincide with the
entire, Euclidean space.
(2) The Principle of Indeterminacy, or Uncertainty, which W. Heisenberg
succeeded in making certain and then,in cooperation with N. Bohr, broadened
into the comprehensive philosophy called the Copenhagen Interpretation, is
clearly anti-realist. In spite of this geographical limitation, the anti-realist
implications of this philosophy are pervasive. Obviously, no geographical
limitations apply to my attempt at showing the fallaciousness of the anti-realist
argument from the Indeterminacy Principle.
(3) In his attempt to provide a solid, mathematical restructuring of non-
relativistic quantum mechanics, von Neumann did not espouse the Copenhagen
Interpretation but succeeded in proving an intriguing, epistemological theorem,
dealing with the arbitrariness of shifting the 'cut' separating the measured object
from the measuring conglomerate. The conglomerate is usually con-strued as
involving the activity of a conscious, human observer. This is why von
Neumann's theorem is often claimed to have anti-realist implications. This claim
is disclaimed in my discussion, but the ability of 'proving' an epistemological
level is nevertheless credited to von Neumann. It may be the
INTRODUCTION 21

case that only he could have succeeded in making a demonstrably valid,


epistemological fmding.
(4) The new role of probabilistic concepts in all quantal theories has often
been viewed as incompatible with philosophical realism. The use of the calculus
of probability in empirical endeavors is not new, since four centuries ago, Blaise
Pascal had invented it, in order to help an unlucky friend, who may
have been a compulsive gambler (Chevalier de Mere). The calculus, reasonably
established by the Swiss family of the Bemoullis, has kept invading strategic,
scientific theories, like genetics, thermodynamics and statistical mechanics.
But the quantal role of the probabilistic calculus has been new in a clear,
non-controversial way. The novelty was often construed as incompatible
with epistemological and/or ontological realism.
(5) For the first time, in the history of empirical sciences, unorthodox, new
logical formalisms have been associated with non-relativistic quantum
mechanics. These new formalisms were rooted in non-classical versions of the
most fundamental logical theory, viz. the propositional calculus. Dutch neo-
intuitionists Brouwer and Heyting have already tried these innovations in their
attempts to obtain, a more secure foundation for the assembly of all
mathematical theories. They apparently felt that the 'logicist' foundations,
credited to Frege, Russell and Whitehead, were inadequate. They also rejected
the 'formalist' foundations established by Bemays and Hilbert. The Dutch
mathematicians might have claimed the somewhat earlier support of the Russian
mathematician, Kolmogorov (whose paper, published in Russian,
was discovered much later by an American, A. Church, who knew more than
anyone else in this area). Yet, an unorthodox logical formalism, allegedly
fitting into or required by an empirical theory, was suggested for the first time in
history by von Neumann (again, his name). An outline of his un-orthodox,
propositional calculus, is to be found in three or four pages of his 1932 treatise
[Mathematische Grundlagen der Quantenmechanik]. It was then elaborated
upon in a joint effort of G. Birkhoff and von Neumann and continued by many
scholars interested in the foundations of quantal theories (e.g., Paulette Fevrier,
Hans Reichenbach, Joseph Jauch and J. Zeman). If an 'empirical logic' were
required to make sense of the empirical quantal theories, then an anti-realist
argument would certainly become cogent.
Needless to say, the strategic importance of quantal theories in contem-porary
science more than deserves an autonomous philosophical investigation, rather
than the subsidiary role allotted to its relevance to the joint problems of Time and
Causality. In this restricted context, I have tried to avoid some risky topics, even
if they are potentially relevant: thus, I have left the treatment of
22 INTRODUCTION

the sub-particles called 'quarks' to some younger scholars, who may live long
enough to afford a safe treatment of these intriguing entities which have
apparently spread from Pasadena, California, to several countries, on both
sides of the Iron Curtain (and, possibly, on the tricky top of the Curtain,
where some 'non-aligned' scholars might be found). Nobody can doubt the
size of the philosophical impact of quantum science, if he just reads the
inscription, "The Lord does not play dice," engraved at the Princeton Institute
for Advanced Study suggested by Albert Einstein. In this case, again, we
have the advantage of hindsight, and we realize now that his theological
statement is wrong. But Einstein's mistakes are still more important than the
truths discovered by other scholars. (Thus, those who agree with his
inscription had to go into hiding and make the 'hidden variables' responsible
for the allegedly misleading appearance of the breakdown of determinism.
The hiding device recalls several respectable names, e.g., Bohm and Vigier.)
Actually, nothing is more comforting for people like me, who dare to
comment on the views of Einstein and many other leading scientists who are
no longer with us, like N. Bohr, W. Pauli or M. Planck - than the other,
Einsteinian inscription, engraved at the Institute: "The Lord is subtle, but He
is not mischievous." I humbly hope that the Divine tolerance is granted not
only to scientists who have kept discovering some tricky, natural laws, but
also to my fellow philosophers who try, to the best of their ability, to solve
the somewhat less tricky problem, dealing with the meaning, the scope, and
the substantiation of the natural laws discovered by the natural scientists.
An alternative axiomatization of the relativistic theory of space-time is
outlined in 'Relativity and the Atom'. The only undefined term is 'collision-
connectibility' interpreted as the possibility that two world-points A and B may
be the non-simultaneous locations of one and the same particle such that another
particle could have collided with the former at both world-points
A and B. 'The possibility of a collision of some particle C with a particle D'
is construed as the valid claim that no natural law would be violated if such a
collision should actually occur. The choice of a symmetrical relation between
world points makes it possible to defme a generalized, inertial frame of
reference in terms which are meaningful at the three levels of spatio-
temporal extension (microphysical, macro physical and cosmological). The
symmetry of the relation is chosen in order to avoid any commitment about
the symmetry, or parity of universal time.
The detailed comments on the proposed axiomatization establish the
derivability of the Lorentz transformation group for spatio-temporal co-
ordinates without raising the objection to which several extant axiomatic
INTRODUCTION 23

systems tending to ensure the validity of the Lorentz group are open. Such
derivations have been attempted by Einstein himself, and then by Broad,
Caratheodory, Pauli, Fock, Robb, Stiegler, von Ignatovsky, Frank and Rothe, and
Reichenbach. The puzzling feature of most of these attempts is their endeavor to
derive the Lorentz group from the laws governing the propagation of light, i.e., at
bottom, from Maxwell's electro-magnetic equations. Yet, the impossibility of
such a derivation is shown by a well-known, group-theoretical finding, often
called 'Bateman's Theorem', although not only Bateman but also Cunningham
was involved. This 1910 theorem shows that Maxwell's equations are invariant
under two groups of transformations, viz. the linear Lorentz group and another,
non-linear group. H. Weyl's 1925 review of H. Reichenbach's monograph,
published in 1924, shows exactly the intrinsic limitations of a purely
electrodynamic approach to the problem ofaxio-matizing the theory of space-
time, inherent in Einstein's Special Relativity and somewhat re-formulated by H.
Minkowski who preferred to resort to the language of pure mathematics. Weyl's
critical remark did not keep the group of the aforementioned scholars from
relying on the electro-magnetic approach. Thus, even C. Caratheodory, although
very familiar with Bateman's Theorem, was moved by some reasons of his own,
to espouse the electro-magnetic approach - to the delight of Albert Einstein. This
physicist may have been closely associated with H. Reichenbach, but he
preferred to present just the Caratheodory result to the Prussian Academy of
Sciences. Both Caratheodory and Reichenbach published their findings in 1924.
Obviously, the history of the sciences, and of their philosophies, is as misleading
and twisted as is, for example, the political history of socialist systems (which
depends not only upon what actually happened in the past, but also upon what
happens to be the present geographical location of the historian).

Several other issues deal with the relevance of the TCP theorem to tem-poral
symmetry, the pretty integral, rather than fragmentary picture of physical reality,
provided by Special Relativity (in particular, all the Lorentz-invariants, like rest-
mass, electric charge, time-likeness or space-likeness of Minkowski-intervals,
etc., are observer-independent, intrinsic features of phy-sical reality). There are
also meaningful restrictions on the quantal breakdown of strict determinism,
since this breakdown only affects the deterministic nature of some physical
processes, but retains the deterministic, exactly pre-dictable nature of an
extensive class of other processes. Thus, non-relativistic quantum mechanics
provides for a rigorous prediction of any future quantum state of a closed system
whose present state has been determined by successful measurements of a
complete set of commuting operators, representing an
24 INTRODUCTION

appropriate set of measurable quantities. There is another, puzzling way of


speaking used in current physical theories, viz. the talk about 'virtual' pro-
cesses in which the conservation laws are suspended. To my mind, there is
little theoretical virtue in the virtual talk. In older theories, the adjective
'virtual' was sometimes used as an auxiliary way of referring, in variational
principles of many physical theories (e.g., the Principle of Least Action), to
'virtual displacements', etc. At a later stage, rigorous mathematical for-
mulations of variational principles have been made available owing to the
emergence and growth of Functional Analysis. Hopefully, a similar, non-
problematic replacement will be found for other references to 'virtual'
processes.
Chapters IV and V of Part III deal with the direction, or arrow of time - or
correspondingly, with the symmetry and parity of time. This problem has already
been partially explored in Chapter 7 of Volume I. Yet further treat-ment is
imperative, both because of the relevance of the many fmdings I have just
mentioned and of many others, which are equally relevant. Thus, the problem of
the reversibility of organic evolution is scientifically and phi-losophically
important. The question is whether our human, tripartite classi-fication of all
mental and physical events into those which belong to the past, or those that are
included in some reasonably extended (not pointlike) present, or those which will
occur in the future, is based on an objective, observer-independent significance of
the idea of temporal succession? This would be the case if time were anisotropic.
But should time be isotropic, symmetrical, then we would have to assume that the
evolutionary process of terrestrial life was somehow started, perhaps 2 to 3
billion years ago, b'y some accidental emergence of primitive life caused either
by terrestrial, chemical processes, or by the impinging of some extra-terrestrial
living entities at an early phase of our planetary history. The question of the
origin of terrestrial life does not matter here. It may well be that some
atmospheric condition prevailing at this early stage of terrestrial history (perhaps
just the kind of condition once assumed by Urey) was conducive to the
emergence of life on earth. Our problem is whether the laws governing terrestrial,
organic evolution are time-reversible. In other words, we are asking whether an
initial stage of life, on some other heavenly body similar to the terrestrial stage,
but involving
a reversed temporal orientation, would have resulted in a lengthy, evolutionary
process temporally oriented not from our past toward our future, but from their
past (our future) toward their future (our past)? The question looks very
speculative, but, nevertheless, it does admit a simple answer. The single decisive
factor which brought about a lengthy accumulation of very small
INTRODUCTION 25

changes caused by natural selection and the survival of the fittest was dis-
covered by the Dutch scientist, de Vries. He found that 'mutations'
correspond to Darwin's small changes. Some decades later the lengths of
Darwinian evolution were determined by two British scientists, Sir R. A.
Fisher and J. B. S. Haldane, who computed the duration of the evolutionary
process in terms of 2 billion to 3 billion years. The problem of evolutionary
reversibility amounts to the reversibility of the 'mutations'. We now know that
the molecules called genes perform one single, crucial function, viz. they
transmit genetic information. And the theory of information which. also
originated just a few decades ago, has now come of age: information turned out
to be negative entropy. And contrary to the alleged irreversibility of entropic
changes, which has misled people like A. Eddington (and is still misleading
many believers in the 'Second Principle of Thermodynamics') it was refuted in
the second decade of our century by P. and T. Ehrenfest and M. von
SmoluchowskL Th(ly did establish the cyclical, rather than the irreversible,
nature of. entropic changes. The same biological laws would apply to this
hypothetical, evolutionary process as they do to ours, which is no more
hypothetical.
Yet, the problem of evolutionary reversibility significantly accounts for just
one puzzling feature of time-symmetry. In surveying the principal theories of
physical science, we shall realize, from the outset, that this problem is not self-
contained. Thus, already in the case of Newtonian mechanics expressible in the
Laws of Motion, the second Law, which equates the force acting on a body B
with the product of the mass and the acceleration of B, the force may undergo
irreversible changes in time. Hence, solutions of Newton's equations which
violate the symmetry of time can be and have been found. The point is that there
is no evidence supporting the existence of such troublesome forces. When we
proceed from Newton to Maxwell, the reversal of temporal duration (Le., the
substitution of -t for +t) has to be associated with changes of the direction of
some electromagnetic magnitudes, in order to show the time-symmetrical nature
of the Maxwellian Theory. Then comes the turn of thermodynamics, either
phenomenological or statistical. After many decades of misconceptions, both
versions turned out to be time-symmetrical. The eventual result was obtained
only through the co-operation of many leading investigators. Boltzmann (and his
H-theorem) were involved, in addition to Gibbs, Zermelo, CaratModory and
Khinchin (to name just a few). The fact that mathematicians of the caliber of
CaratModory, Khinchin and Zermelo were involved is significant because it
illustrates the mathematical complexity of the issue. Eddington's
misunderstanding of the situation is not significant,
26 INTRODUCTION

but very puzzling indeed. Perhaps as puzzling as E. Mach's inability to put up


with Einstein's Special Relativity, although Einstein's breakthrough might not
have occurred, had the conceptual analyses of Mach's Science of Mechanics not
been available to him. No less a person than H. Weyl has stated that the
irreversible propagation of light emitted by a point-like source provides time
with an arrow. Yet Weyl, whose mathematical mastery of relativity was second to
none (including Eddington's mathematical presentation of Einstein's Theory),
was apparently unaware of the fact that Einstein himself established the
reversibility of electromagnetic radiations generated by a point-like source as
early as 1910 (viz. by using his corpuscular picture of light).
The emergence of quantal theories has started 'a new beginning' for the
overall scientific outlook, including the problem of time-symmetry. The
'beginning' of this branch of science started with the beginning of our century.
No end is in sight. Nor was the nature of time the main question. The most
fundamental epistemological and ontological insights of the sum total of
empirical sciences had been challenged: the role of the conscious observer,
the observer-independent existence and nature of the universe, the meaning of
physical reality, the validity of the spatio-temporal perspective - all these age-
long insights have been affected by the never-ending sequence of increasingly
sophisticated, quantal theories. All I could do in the chapters dealing with the
'arrow of time' is to critically survey the most significant, extant contribu-
tions, not neglecting the early beginnings of quantum physics, but rather
emphasizing the more recent investigations, including some papers published
in 1976. Over the decades, the quantal theories kept multiplying, moving
from a non-relativistic to a relativistic stage, from first quantization to second
quantization, from low energy physics to high energy physics, from cosmo-
logical or astrophysical quantal theories, to those dealing with standard size,
macro physical objects which somehow behave in quantal ways, e.g., in cases
of superconductivity, to molecules, to atoms, to elementary particles. Al-
though there is now some promise of moving from elementary particles to
sub-particles, I could not afford moving to this promised land. I shall not list
here the names of the many investigators, scientists or philosophers of
sciences who contributed significantly to the central issue. I do attempt to
formulate a tentative solution and to present a reasonable amount of support-
ing evidence. It would be utterly unrealistic to expect anything 'final' in the
temporal flux of relevant research and publication.
The features, dealt with consecutively, in Chapters I through VI of Part IV
refer to the measurement of time, to its ontological status, its reality, its
causal nature, its symmetry, and its psychology. In dealing with temporal
INTRODUCTION 27

measurement, I had to refer to the subsequent discussion of von Neumann's


pioneering theory of quantum measurements within non-relativistic quantum
mechanics including both its inadequacy and successful evolution in later
quantum theories. The most essential features of temporal measurement include
the relevance of causal considerations, both deterministic and in-deterministic, to
any measurement of time. Similarly, the interdependence of temporal and spatial
measurements and the impact of Special Relativity can be taken care of at this
juncture. This applies also to the epistemologically crucial aspect of the
measurement of any magnitude which is capable of continuous variations and,
therefore, involves idealizations. In pre-relativistic, Newtonian time, it was
construed as a one-dimensional continuum consisting of point-like instants which
are unobservable in principle. In relativistic, Einsteinian time, merged with the
Euclidean, three-dimensional space into a four-dimensional pseudo-Euclidean
continuum called 'Minkowski's world', the point-like instants are replaced with
world-points, which are still unobservable in principle. To some extent, the pre-
relativistic instants have survived the relativistic merger, viz. in what came to be
called 'proper time'. In any event, the essential unobservability of the temporal
idealizations, regardless of whether they are construed as instants or world-
points, remains and, obvious-ly, entails the essential, empirical unverifiability of
any non-tautological statements about those extensionless, idealized entities.

This raises a serious difficulty from the standpoint of the 'Verifiability Theory
of Meaning', once formulated by the leading American proponents of the
pragmatist philosophy and then miraculously rediscovered by a Viennese thinker,
Ludwig Wittgenstein, the late author of a still most influential treatise
[Tractatus Logico-Philosophicus] , consisting, in the main, of just six
proposi-tions (in addition to several comments on some of these propositions and
some more extra comments on these primary comments). In an earlier work of
mine I have tried to show the untenability of the pragmatist and/or neo-positivist
theory of meaning. But, in the context of Part N, I am trying to take care of this
difficulty by proposing an alternative, 'fmitist' approach to the problem of
measurement of continuous quantities.
As a matter of fact, an attempt to overcome the epistemological difficulties
inherent in the scientific use of idealizations was already made many decades ago
by a most prominent thinker, who, once upon a time, made the heroic move from
Cambridge, England to Cambridge, Massachusetts. In the early twenties A. N.
Whitehead published a few volumes, outlining his 'Method of Extensive
Abstraction', which he probably intended to become a sequel to the monumental
Principia Mathematica which he wrote with Bertrand Russell in
28 INTRODUCTION

1910-1912. As usual, a clear account of Whitehead's method is to be found in a


later work, published by another scholar, C. D. Broad, Scientific Thought, of
1927.
The need for replacing the Whiteheadian approach with an alternative,
fmitist approach can easily be shown: Whitehead has simply replaced the
classical idealizations of unobservable entities, like instants or world-points,
with some neo·classical idealizations, mainly the infmite sets of infmite
sequences of fmite time-intervals, or spatio-temporal regions, on the under-
standing that, from the classical point of view, these sequences would be said
to converge to instants, or world-points. There is no vicious circle in
Whitehead's neo-classical Method of Extensive Abstraction. And the circum-
stance that he favored an 'event ontology' by promoting the class of all events
to the rank of the scientific universe of discourse is not decisive because it is
possible to reformulate this ontology in terms of a spatio-temporal universe
of discourse. The trouble is that Whitehead replaced a set of older
idealizations with a set of somewhat younger idealizations. Granted, age is
just a number, as my younger friends often tell me. But, the 'younger' age of
Whitehead's idealizations does not rid them of the predicament of being
idealizations. And idealizations are suspicious, within and without science.
The fmitist approach to the measurement of time, outlined in Chapter I of
Part IV is just less suspicious.
The subject matter of Chapter II of Part IV is the ontological status of
time. I prefer the adjective 'ontological' to the pretty synonymous adjective
'metaphysical' because the latter has somehow got a derogatory connotation.
And the main claim of Chapter II is much more heavily dependent upon the
contents of Volume II than is Chapter I. The point is I am taking seriously an
idea that originated with a young British gentleman who, once upon a time,
wrote a very troublesome letter to an older, German colleague of his, G.
Frege. The British gentleman, Bertrand Russell, subsequently discovered the
'ramified theory of logical types' in order to overcome the difficulty conveyed
in his letter to Frege. When I admit taking the logical theory of types
seriously, I am stating my commitment to the correspondence-theory, and not
only regarding the meaning of truth, which was shown by Tarski to be
roughly synonymous with correspondence to facts (or states of affairs). The
logical theory of types as applied to any language L implies a stratification of
the expressions of L. A correspondence-view of the theory of types comes to
the claim that expressions of various types correspond to (or denote) entities
of various types. Accordingly, a stratification of denotational expres-sions
implies a stratification of the entities they denote (or of their referents).
INTRODUCTION 29

There is no need, at present, to bother with the uncomfortably ramified type-


theory created by Russell. It was then 'simplified' both by his fellow English-man
Ramsey, and both independently and similarly, by my countryman, Chwistek: in
turn, the Chwistek-Ramsey theory of types has been affected by N. Wiener's
fmding that relations can be defined in terms of classes. As already mentioned,
this enabled both GOdel and Tarski to further simplify the theory of logical types.
Their type-theory has been somewhat improved by Quine, who used their
accumulated fmdings to outline a 'cumulative theory of types'. Some type-free
formalisms not afflicted with Russellian antinomies have also been elaborated.
However, freedom is always expensive, and the price of type-free formalisms
turned out to be exorbitant. The cumulative improvement is not relevant in the
context of my investigation.
Hence, the ontological status of time, as construed in the framework of the
Godel-Tarski theory of types, coincides with the rank of time in their hierarchy of
classes. Some philosophers have claimed that assigning a rank in this hierarchy
to any set of entities is a matter of taste. Thus, natural numbers have rank 1 in the
Bernays-Hilbert approach to foundations of mathematics, but their rank is 3
rather than 1 in the Russell-Whitehead treatment of the same subject. The
arbitrariness or freedom of choice has been called 'ontological relativity' by
Quine. In various chapters of Volume II, I show that there is no such freedom in
the empirical area of the physical sciences and that the role of individuals has to
be assigned exclusively to all the observable denizens of Minkowski's world, i.e.,
to all fmite spatio-temporal regions.

In psychology, and other studies dealing with mental entities, the role of time
is obvious, but the role of space is dubious. However, if we disregard the non-
existent studies of disembodied minds, it is reasonable to associate with the
temporal dimensions of the mental entities of any person, the simultaneous
spatial location of his (or her) body, or brain, or a more specific part of the brain
somehow correlated with the mental event. In this somewhat artificial way, the
spatio-temporal universe of discourse could accommodate all empirical sciences,
physical, Psychological or mixed. If someone dislikes the artificiality of a spatio-
temporal universe of discourse for psychological talk, then he can single out what
a physicist calls 'proper time' and promote it to the rank of specific universe of
discourse of psychological talk. Little is gained by this device.

The reasons for identifying the universe of discourse of any language capable
of expressing all available empirical sciences are complex and numer-ous. They
are listed and outlined Part IV of Volume II. However, the main
30 INTRODUCTION

argument supporting my advocacy of the spatio-temporal monopoly can be


stated simply. Ultimately, all the empirical evidence supporting any
scientific theory, however technical and sophisticated, must be expressed in
some 'ordinary language' with a universe of discourse common to this
language and to all presently available scientific theories. On closer
examination, only the spatio-temporal universe of discourse satisfies the
requirement. For example, in General Relativity, the universe of discourse is
clearly spatio-temporal, since material particles are treated as singularities of
the geodesic lines in some non-Euclidean space-time. In Chapter III of Part
IV, the reality of time is explored. But the gist of the claim of temporal reality
can be stated in a few sentences. Only objects which exist in the logical sense
can be real. The logical sense of existence is explicated in an elementary part
of logic: the predicate-calculus, or the quantificational calculus, which
presupposes only the propositional calculus. 'There exists an object with the
property P' can be construed as equivalent to the statement: 'It is not the case
that every object lacks the property P'. Not all existing objects are real.
Natural numbers demonstrably exist, but they are not real. To be real, an
object has both to exist and to be observable in principle. Here 'X is
observable in principle' means that no natural law would be violated if X
happened to be observed. And, to 'observe an object X' means 'either to
manage to perceive' X (in order to· answer some question about X) without
the aid of an instrument (e.g., a microscope or a telescope) or to measure it,
or to obtain reliable information about X by validly computing the results of
measurements performed on X possibly with the aid of some observational
instruments which, in addition to the measuring devices, broaden the scope of
sensory perception. Hence, temporal intervals (or, preferably) proper-time
intervals) exist really because they are effectively measurable and each of
them has obviously one funda-mental property, viz. a specific length. In other
words, time-intervals whose length is I second exist because it is not the case
that everything lacks the property of lasting for 1 second.
The next, fourth Chapter of Part IV deals with the causal nature of time. Here
again, use is made of the many implications of quantum physics regarding the
meaning and scope of causality. This was discussed in detail earlier in Volume II
and the possibility of commenting on the causal nature of time in this chapter
comes from the modification of the deterministic causality which the emergence
of quantum physics has made necessary. I have suggested reformulating this
principle as that of indeterministic causality and to derive the theory of universal
time from the modified causality principle.
An elementary example of indeterministic causality is provided by
INTRODUCTION 31

perceivable changes brought about by measurable decreases of temperature.


Thus, if a glassful of water is cooled below the freezing point, it will freeze in
most cases, and the drop of temperature T will be considered as the cause of the
freeze F. However, the water may be 'supercooled' and fail to freeze in spite of
the temperature drop. Hence, T is only an indeterministic cause of the freezing T
because the occurrence of T raises the probability of the subsequent occurrence
of F, but may fail to bring F about. In other words, the earlier event E is an
indeterministic cause of the later event E' if, and only if, the earlier occurrence of
E raises the probability of the subsequent occur-rence of E'.

Then, in Chapter V, a discussion of the symmetry of time is outlined. To


obtain a reasonably complete account of the main aspects of the problem of time,
it is necessary to discuss its symmetry in a preparatory way.
Roughly speaking, the claim of temporal symmetry amounts to denying any
intrinsic difference between past and future. Heraclitus once said that "the way up
and the way down is the same way." The symmetry of time can be expressed in
this Greek-Ionian mode of speech by denying that time is a one-way street. In the
technical slang of the science and the philosophy of physics, the symmetry of
time is also referred to as the 'isotropy oftime' or the 'invariance of natural laws
under time reversal'. Both the scientific and the philosophical implications of
temporal symmetry are considerable. Thus, if there is no intrinsic ('law-like' or
'nomological') difference between past and future, then all those religious
outlooks which involve the creation of the universe by a Supreme Being at some
point in the past would have to be readjusted to the lack of an intrinsic difference
between past and future. There is no point in questioning the ability of creationist
religions to make the relevant readjustment: in the past, they have already
managed to adjust to the Darwinian theory of organic evolution. The present case
of time-symmetry is hardly more difficult than the evolutionary case. The
philosophi-cally important issue is the need for a new readjustment, rather than
its feasibility.

No wonder that the bibliography of investigations into the symmetry of time


(or, in Eddington's parlance, the lack of an 'arrow of time') is sizable, and,
sometimes obsolete. Since Clausius claimed in the last century that the energy of
the universe neither increases nor decreases, and until the current controversy
about the alleged violations of time-symmetry by recent findings in elementary
particle-physics, a legion of scholars kept investigating the issue. In 1976, Robert
G. Sachs graciously mailed me two of his reprints dealing with the alleged
violations of temporal symmetry by the nasty behavior of
32 INTRODUCTION

Ko mesons which decay into a couple of pions in an allegedly irreversible way.


The complex and ramified issue of temporal symmetry is discussed in several
sections of my work. At this juncture, it may suffice to mention one avenue of
approach outlined in the last section of Part IV. The symmetry of time depends
upon the nature of universal, physical laws, rather than on the occurrence of
particular facts. Thus, even if the 'big bang' cosmological theory were valid, and
the age of the universe could be approximately determined by referring to
Hubble's constant or any other device, this age would still constitute a single
fact, and not exemplify a universal law. If I may use a terminology traceable to
the German philosopher Heinrich Rickert, then I would say that the symmetry of
time is a 'nomological', rather than an 'idiographic'issue.

At present, virtually all reasonably established physical theories consist of


sets of ordinary or partial differential equations involving four spatio-temporal
co-ordinates to solve these equations. The equations of General Relativity are just
a case in point. These equations are known to be invariant under time-reversal,
i.e., roughly speaking, the substitution for the time variable t of -t does not
change the equation. Actually, some other changes are also involved and their
nature depends upon the nature of the particular theory under consideration. This
situation (I mean, the nomological irrelevance of the direction of time) was stated
concisely in The Character of Physical Law published by R. P. Feynman in
1965. But, in the ensuing decade, many relevant fmdings were made and a pretty
confusing situation came about. Some easygoing researchers may feel that the
situation is hopeless, but not serious. I, for one, feel pretty strongly that the
situation is serious, but not hopeless. In any event, an important aspect of the
issue depends not only on the fundamental equations but also on the kind of
solutions which can be found for them. Solving a system of partial differential
equations, e.g., those of Einstein's General Theory of Relativity, has always been
a most difficult job, even if an outstanding mathematician's book, How to Solve
It [G. Polya] is made use of. The mathematically sophisticated physicist has to
assume some reasonable conditions which are fact-like rather than law-like
(idiographic, rather than nomological). Depending upon the mathematical nature
of these fundamental equations, the conditions required for their solutions may be
of an 'initial' type, or a 'boundary' type or, sometimes., of an 'asymptotic' type
(e.g., referring to conditions prevailing at t = too). My reference to the human
factor involved in obtaining a valid solution to a set of fundamental equations is
oversimplified. I have mentioned the mathematically sophisticated physicists as
the people most likely to successfully handle the challenge. Perhaps I
INTRODUCTION 33

should also have mentioned the fact that a most prominent mathematician,
D. Hilbert, has once stated that physics is too difficult for a physicist, only a
mathematician should deal with the relevant mathematical problems, and the
principal exponent of General Relativity, H. Weyl, an equally outstanding
mathematician, did not try to integrate Einstein's partial differential equa-
tions. And an unsurpassed mathematician, J. von Neumann, was so baffled
by the intricacies involved in solving partial differential equations that he
recommended resorting in such cases to a demonstrably super-human, prob-
lem-solving device, viz. a computer whose controlling variable is capable of
a continuous variation (Le., an analog, rather than a digital computer).
Fortunately, solving systems of partial differential equations is a human
task. And the decisive fact is that the successful scientist faced with such a
problem can do without an analog computer and need not be either a
physicist or a mathematician; sometimes, it suffices that he be a man of
genius. This was the case of K. GOdel, who succeeded in solving the partial
differential equations of General Relativity in 1949. His finding started a
bizarre sequence of scholarly publications whose outline will be presented in
this context, because this may help to prevent future occurrences of similar
sequences.
(1) In his search for cosmological models which will exhibit rotation, Godel
discovered a metric satisfying Einstein's equations with a non-zero cosmo-
logical constant. Godel enumerated many remarkable properties of his solution
including the fact that his metric allowed closed, time-like curves. This last fact
implies that a traveler in a spaceship, expending a sufficient amount of energy to
travel around the universe, can return to his past, but such travel
would take a time comparable to the age of the universe and require a corre-
sponding amount of energy. However, in his paper, Godel did not state
whether he considered freely-falling observers and whether they can
describe time-like curves (which would be geodesics in his case).
(2) In a paper published in the Proceedings of the National Academy of
Sciences in 1961, S. Chandrasekhar and James P. Wright explicitly integrated
the time-like geodesics in Godel's metric and showed that there are no such
closed curves. In fact, the paper admits the existence of closed, time-like
curves in the Go delian universe but shows that such curves are not
geodesics (i.e. shortest, or, preferably, extremal curves in non-Euclidean
spaces). The Significance of the failure of a curve to be a geodesic line is due
to the circum-stance that in. General Relativity (axiomatically in Einstein's
early papers, and, demonstrably, on the basis of Einsteinian equations,
according to a result obtained by Fock at a somewhat later date), material
particles are
34 INTRODUCTION

construed as singularities on geodesics. Hence, no spaceship could possibly


travel along a closed, time-like curve in a Godelian universe. Thus, Godel's
troublesome paradox vanishes altogether. In other words, they showed that
freely-falling observers (that is, observers not expending any energy) cannot
return to their past. James P. Wright told S. Chandrasekhar that he had been
informed by Godel (in an interview) that the Princetonian had considered this
matter and arrived at the same conclusion that the authors of the 1961 paper
had reached.
(3) At a later symposium on the nature of time arranged by T. Gold (and
summarized in the volume, The Nature of Time, edited by him in 1967), S.
Chandrasekhar is reported to have stated that he had shown that GOdel's
metric did not allow closed time-like curves. But Chandrasekhar has
informed me recently, in addition to what I stated in the preceding paragraph,
that the report is inaccurate and that it mis-states what he had said; and that
his report at the meeting was only a summary of his published paper.

The alternative solution of the paradox ascribed by Gold to Chandrasekhar is


due to the fact that Chandrasekhar had no opportunity to check on the chapter
reporting on his address delivered at the aforementioned Symposium. It is
somewhat distressing (for those who may be inclined to ascribe a super-human
stature to Einstein, instead of a simply unique supremacy in our century's physics)
that Chandrasekhar, who sat in at the Princeton meeting when Godel presented his
rmding to Einstein, that Einstein 'disliked' the
rmding but raised no theoretical objection to it. This simply instantiates the
fact of man's fallibility and parallels the congratulatory letter once addressed
by Einstein to Schrodingerin connection with the latter's version of quantum
mechanics in conjunction with Einsteinian reservations about the Heisenberg
version. The equivalence of both versions was established by Schrodinger at
a somewhat later date.
(4) Based on Gold's report, H. Stein wrote a rejoinder confirming Godel's
original statements as against Chandrasekhar's. (Cf. H. Stein: 'On the
Paradoxical Time-Structures of GOdel', Philosophy of Science 37, No.4,
1970.)
(5) The practical conclusion derivable from the aforementioned sequence
of publications is the need for a yearly publication containing abstracts of all
papers published in the preceding year and dealing with topics overlapping the
research areas of astronomers, cosmologists, logicians and philosophers of
science. This would have prevented T. Gold from summarizing Chandrase-khar's
views in an inaccurate way and saved H. Stein the effort involved in
INTRODUCTION 35

writing his remarkable, but unnecessary comment on the alleged Godel-


Chandrasekhar controversy.
In his recent letter addressed to me, S. Chandrasekhar pointed out the
existence of other, non-GOdelian cosmological models of General Relativity
in which closed, time-like curves are available, but do not overlap with the
class of geodesics. If this should tum out to be the case for all cosmological
models of General Relativity, then the nature of relativistic time may be
affected. Obviously, we have to assign this issue, and most related issues, to
some other generation of scholars.
However, currently the problem of temporal symmetry is mainly affected by
several fmdings in the theory of elementary particles, e.g., Ko mesons or
deuterons. Thus, the decay of Ko mesons into two pions was construed as a
violation of time-parity in 1976 by R. G. Sachs and, in earlier years, by Casella.
On the other hand, the relevance of these fmdings to the parity of time has been
denied by other scientists (e.g., Conforto, Neeman, and Feinberg).
The gist of the defense of temporal symmetry outlined in the penultimate
section of Part IV can be stated at this juncture: only some fmdings in
elemen-tary particle physics are currently viewed by general scholars as
relevant to the nomological irreversibility of certain processes in this area.
These fmdings deal with the behavior of several particular species of
elementary particles, e.g., Ko mesons or deuterons. Thus, the decay of Ko
mesons into two pions fails to provide time with an arrow if no Ko mesons
ever materialize. The availability of these mesons at any particular instant of
cosmic history does not follow from the sum total of universal laws of nature.
We may assume, e.g., that, 'in the beginning' every particle collided with its
antiparticle and the entire population of the universe consisted of photons,
which are ad-mittedly governed by time-symmetrical laws. This assumption
is idiographic, rather than nomological. And so is the additional assumption
that, in this remote past, there was a distribution of occupation-number
operators which, through successive pair creations, would have eventually
brought about the present population of all kinds of elementary particles in
the entire universe. The conclusion entailed by these highly speculative, but
still admissible assumptions amounts to the following: processes violating the
symmetry of time occur only locally. Universal time has no arrow. In other
words, the laws implying the irreversibility of certain processes may be valid.
But, throughout time, they are only vacuously valid, since the processes
described by these laws may fail to occur.
This is just one promising attempt to get rid of the arrow of time. Several
other attempts, perhaps more promising, have been made by several scholars
36 INTRODUCTION

and are discussed in Volume 1. Needless to say, all these attempts are
tentative, and so are the attempts to save the arrow of time. The issue is
momentous, like many other issues in philosophy, in science, and in the
overall issue of mankind's survival. It stands to reason that any solution to
any momentous problem can only be tentative.
ESSAY ON THE CAUSAL THEORY OF TIME

PART ONE

THE CAUSAL THEORY OF TIME IN THE WORKS OF ITS


PRINCIPAL REPRESENTATIVES
INTRODUCTION TO THE ESSAY

In the present study we propose to recount the evolution of the causal theory
of time in its broadest outlines. This theory is meant to reverse the classical
explanation, which deduced the relation of cause and effect from the tem-
poral relation of before and after. Even those who disagreed with Hume's
assertion that causality could be reduced to a regular succession in time -
contending that causality involves a dynamic link not necessarily present in
certain regular successions (such as Schopenhauer's example of days and
nights) - nevertheless considered the temporal order of succession as the
fundamental order on which a special dynamic link could be superimposed
in certain cases, thereby constituting the causal relation. They shared Hume's
opinion that there can be successions without causality and even saw here
the normal sort of succession. If science taught them that universal
interaction made succession without causality impossible, they saw this an
empirical complication which would emich the pure temporal relation, which
itself could exist a priori, without any dynamic link.
The causal theory of time maintains the contrary. It considers the dynamic
causal order of becoming as the fundamental fact, from which the temporal order
of succession, simultaneity, and duration is deduced as a simple conse-quence. If
event X takes place before event Y, it is because X has contributed to the
production of Y. In the classical explanation, X contributed to the production of Y
because Y followed it regularly, that is, in conformity with a causal law . But this
theory contains a vicious circle according to the causal theory of time, for which
the temporal order of succession is nothing but the simplest outline of the causal
relation. The same is true of simultaneity: in classical theory there can be no
causal action between two simultaneous events provided that the principle of
[continuous] action, point to point, is taken for granted, excluding all
instantaneous propagation. According to the causal theory of time, two events are
simultaneous by definition if there can be no causal action between them
(furthermore, this is just one of the possible causal defmitions of Simultaneity).
The same is also true of the "third mode of time," to use Kant's term, that is,
duration: in the series of successive states of a substance enduring in time, (for
example, a material particle) every state plays a preponderant role in the set of
conditions determining its
39
40 THE CAUSAL THEORY OF TIME

successive state; the influence of the states on other substances comes into
play only in circumstances of second order (we will make these vague
notions of 'preponderant role' and 'circumstances of second order' more
precise in what follows).
Here again the causal theory of time reverses the classical explanation: in
order for a series of events to be considered by definition as the succession of
states of a substance which endures in time, it is necessary and sufficient that
its elements be arranged so that each one 'plays a preponderant role' in the set
of conditions determining the following element. The disposition of events in
the order of before and after, the analysis of becoming into 'cross-sections of
simultaneity', its 'fibrillous' structure determined by the direction of
'substantial lines' in the Minkowskian universe - in short, all the properties of
time, considered as an order, would be defmable as a function of the causal
relation. But the same seems to be true of time considered as an extended
magnitude. We know that, in the final analysis, every temporal measurement
amounts to the comparison of two intervals of time. Following the usual de-
finition, two intervals of time are of equal magnitude if they are respectively
simultaneous with two periods of a periodic process traversed by an isolated
material system. But the most general definition that can be given for an
isolated system is that in such a system each instantaneous state contains the
ensemble of determining conditions, i.e., the cause of the following state.
Thus the notion of the equality of two time intervals, which is sufficient to
define every metric of time, can be deduced in turn from the causal relation.
From this we can conclude that the metrical properties of time, just like its
ordinal properties, can be defined as a function of the causal relation.
The causal theory of time found valuable support in Einstein's theory of
relativity, and hence recent works (of Robb, Lewin, Reichenbach and Carnap)
devoted to the causal theory, have taken relativity as point of departure. However,
here we are dealing with an epistemological theory which claims not to add new
explanations or facts to those already known to science, but only to delimit the
epistemological significance of scientific data by taking into account the part
played by convention, fact, conceptual construction and intuition. In principle,
the historical study of this theory presents the same difficulties as in the study of
any other doctrine which is integral to the philosophical systems of which it is
part. Thus in the particular case which I have in mind, Leibniz's causal theory is
inseparable from his monadology, and Kant's theory from his transcendental
idealism with all the formidable conceptual and terminological apparatus of the
Critique of Pure Reason. In every attempt to isolate one common theory from
several systems, we clearly
INTRODUCTION 41

run the risk of distorting its spirit, even though one theory may have arisen from
the other, and even if the different formulations, taken literally, are susceptible to
a common interpretation. l I have tried to circumvent this real difficulty by
isolating the causal theories as far as possible from the systems of which they are
part, bearing in mind their differences as well as their common background in the
exposition and critique that follow.
What we are dealing with here, of course, is the causal theory of time
considered as an ensemble of problems and methods, the history of an idea and
not that of its creation. We will not be concerned to establish either the probable
'influences' that thinkers have exercised on one another or the origin of any of
Leibniz's or Kant's ideas; rather we simply want to trace the intrinsic evolution of
the idea in broad outline, that is, the series of causal theories of time in order of
their complexity and the range of the explanations which they furnish. The
chronological order of exposition is not essential to this study and has not been
rigorously observed.
May we say a few words about the aim of this book. We undertook it because
we were attracted not only by the fine example furnished by the causal theory of
time of the close cooperation between philosophy and science, but also by the
interest which this theory arouses in the current state of philosophical discussion
of the problem of time: it is this theory alone, in fact, which seems to furnish a
positive, if not definitive, solution. In addition, more room is given here to
critical analysis than to the historical reconstruc-tion of doctrines.

The reader should excuse the inevitable gaps in a reconstruction attempted for
the first time. As far as criticism is concerned, we have tried always to round it
out with positive remarks, indicating either modifications to be introduced in the
ideas which are criticized or concepts to take their place. However, these
concepts could only be sketched in the course of an historical study: their
systematic development has been reserved for the second part of the present
essay where a new causal theory of time will be presented.
LEIBNIZ AND THE BEGINNINGS OF THE
CAUSAL THEORY OF TIME

1. THE RELATIONAL THEORY

In 1716, the very year of his death, Leibniz, just after expounding his rela-
tional theory of time (usually referred to as his 'relativist' theory) in a famous
exchange of letters with Clarke, rounded it out with a causal theory. The new
theory owed its composition to an outside stimulus, the 'excellent mathe-
matician', Christian Wolff, who had published an article in a recent issue of the
Acta eruditorum in which certain of Leibniz's ideas on mathematical proof were
set forth. Leibniz seized the opportunity to write down some thoughts,
"contemplated for a long time", on the "metaphysical origins of mathematics".
The causal theory of time is found in several lines of the note which is supposed
to defme the 'mathematical' notions of space, time, distance, etc., in terms
borrowed from the 'metaphysical' notions of sufficient reason and non-
contradiction. The theory had no immediate effect. It was not until 1863 that,
thanks to the efforts of Pertz, this note (based on a manuscript in the Royal
Library in Hanover) was printed in the 'complete' edition of Leibniz' works.2 It
was the object of a brief critical exposition in Baumann's compilation [1869] on
the philosophy of mathematics,3 ap-parently without either the exposition or the
critique attracting the attention of those interested in the problem of time. In his
great work [1902] on Leibniz, Cassirer 4 was perhaps the first to grasp the
philosophical interest of the theory and its intimate relationship with the views
which Kant expresses in the 'Transcendental Analytic'. Finally, in an article
published in the Kant-studien [1924] Reichenbach 5 has stressed the strange
coincidences which marry the causal theory of Leibniz with the recent theories of
Carnap and Lewin and his own.

For Leibniz this theory was the result of a long series of efforts devoted to
the enigma of time. We must not forget that it was he who established the
place of the idea of time in the modern mind. In the seventeenth century,
knowledge of philosophical problems relating to time was doubtless quite
summary. Those paragraphs in the fourth book of Aristotle's Physics (whose
arithmetical and essentially purely verbal definition. of time had survived
twenty centuries and influenced the thought of Leibniz himself), and the fine
42
LEIBNIZ AND THE BEGINNINGS 43

dialectic of the eleventh book of St. Augustine's Confessions constituted, by


themselves, almost the whole legacy of antiquity.6 Descartes, whose physics
was based on the identity of substance and extension, failed precisely because
of his misconcpetion of the role of time,7 and added little to this legacy.
Spinoza made thought and extension attributes of substance, without taking.
account of time, which is nevertheless a connecting link between them. Thus
it is from Leibniz that the parallelism of space and time stems, so
characteristic of the entire modern evolution of the problem of time. 8 No
doubt this parallelism, after having been maintained for a long time, became
dangerous in that it tended to conceal the profound problems peculiar to time:
this is why the reaction, begun by Bergson, in favor of a non-spatialized time,
has been necessary, and has found many followers. But side by side with the
divergences are fundamental analogies between space and time which, his-
torically speaking, justify the Leibnizian parallelism. Furthermore, the
modern evolution of the problem of time, after the long domination of
Kantian parallelism, which sees in space and time some forms of sensibility,
actually seems to be reverting to Leibnizian parallelism, for which space and
time are types of order, time in particular being "the order of successive
phenomena" .
The causal theory of time goes further than this relational theory; it does
not stop at affirming that time is just a set of relations, but claims to explain
the nature of temporal relations. Before giving the interpretation of this
defmitive theory of Leibniz, let us briefly recall his relational theory. We
shall see, in fact, that this theory, defended by Leibniz throughout the whole
of his long career, and known only as his 'relativist' doctrine,9 and moreover
seen only through Kantian criticism in the refutation of the Transcendental
Aesthetic, naturally had to lead to the causal theory simply by means of the
application of the general principles of Leibnizian philosophy.
How should that statement to which Leibniz returned so willingly - name-
ly, that time is nothing but the order of successive phenomena - be under-
stood? Certainly not in the precise sense of a relational structure, for which
we now use the term 'order' in pure logic and related sciences. For Leibniz,
who, as partisan of the logic of attribution, did not acknowledge the relations
of reality proper, this dictum must have had (as is already immediately in-
dicated by its form) a negative sense: time is not a substance, as 'the English',
Le., Henry More, Newton, Clarke, believed. For them, time is in some sense
prior to the events which take place in it. It forms a single whole (there is
only one time, as Kant would say) whose ultimate elements, the instants,
occupy, each in relation to the others, absolutely ftxed positions, as to size
44 THE CAUSAL THEORY OF TIME

and sense. The order of events is derived from the order of the instants of the
time which the events occupy: event A precedes event B if the instant
occupied by A is prior to that occupied by B. According to this substantialist
theory, there are no direct temporal relations between events: if the life of
Aristotle preceded that of Kant, it is because the instants occupied by Aris-
totle were prior to those occupied by Kant. It might happen that by some
miracle Aristotle was resurrected some time after the death of Kant; he would
then be at once prior to and posterior to Kant, but nothing can change the
temporal relations of priority, posteriority, and simultaneity between the
instants themselves, and the instants of the fIrst and second lives of Aristotle
would always remain separated by the instants of Kant's life. It is this order
of instants, as prior to, and independent of, the order of events, which
Leibniz refuses to recognize.
For him, the relations of succession take place directly between events
themselves: "instants, considered without the things, are nothing at all and
... they consist only in the successive order of things ... "10 The temporal
order of instants is derived from these direct temporal relations between
events: the hypothetical resurrection could not take place without involving
serious anomalies in the order of instants, which only reflects the linking of
events.
Such is the negative component of the relational theory - the most impor-
tant in the eyes of Leibniz as well as of his adversaries and followers. But
what is its positive component? Precisely what does it mean to affIrm that
time is 'relative', that it is an order and a relation? Every relation presupposes
its terms, every order is an order of something, time insofar as it is order and
relation is relative to the events which it orders, between which it takes
place; it is nothing outside of them. But this is only the negative thesis which
we have just outlined.
There is, however, an important positive component in the Leibnizian
theory: it is relational but not relativistic. I call a theory of time relational if it
makes time consist of relations; I call a theory relativistic if it makes
temporal relations depend on extra-temporal circumstances (a frame of
reference, velocity, etc.). A relational theory can very well not be relativistic,
although a relativistic theory is necessarily relational. Hence for Leibniz time
is only the order of non-simultaneous events; but, for a pair of non-
simultaneous events, succession is an invariable fact determined in a
univocal manner by the two events. Succession in the Leibnizian theory is, in
away, an absolute relation, independent of every convention and every frame
of reference. This distinction between relational and relativistic theories, or
LEIBNIZ AND THE BEGINNINGS 45

better, between absolute and relative relations, is, moreover, of a general order.
For example, resemblance is not an absolute relation, since two objects can be at
once similar and dissimilar, from different points of view (e.g., similar as to color
but not as to form). In the General Theory of Relativity it is the same with
respect to relative motion. Two objects move or are at rest with respect to each
other according to the chosen frame of reference: the theory of motion is thus
relational and relativistic in General Relativity, but it is only relational in Special
Relativity. It is also clear that every relation can be made absolute if one takes
into account the 'frame of reference' implied, in the most general sense of that
term. Thus resemblance becomes absolute if one is precise as to what it must
mean - which amounts to considering it as a relation not of two, but of three
terms, viz., 'the terms' stricto sensu and the 'tertium comparationis'. In the
same way, motion, even in General Relativity, may be considered as an absolute
relation between the two moving objects and the chosen system of reference.

2. THE CAUSAL THEORY

Leibniz's relational theory is not so much a relativistic theory, an abstraction


made of temporal anamolies, as it is a resurrection, in that it establishes the same
temporal order between instants and events as the substantialist theory, the only
subtle difference being that it considers the order of events to be a fundamental
given. ll In order to mark the transition from the relational to the causal theory, let
us recall that Leibniz, as opposed to Descartes, dynami-cally conceived the
nature of phenomena whose type of order is time. Matter, which is an ensemble
of substantial centers of force, fills the universe in a continuous manner (it fills
the interstellar spaces in the form of ether); every phenomenon is a movement;
every cause of movement is a collision, thanks to the continuity of matter; every
substance acts on every other.
In the series of successive states of the universe there is 'equipollence' as far
as the mechanical principles of conservation are concerned. Nevertheless it is the
prior state which contains the reason for the posterior state and not vice versa:
"the present is always pregnant with the future and ... no given state is naturally
explicable except by means of that which has immediately preceded it." 12 In a
letter to Bourguet, Leibniz put forward the following difference between the
points of space and those of time: none of the points of space has any logical
priority over any of the others, but if one instant precedes another, it is prior not
only temporally, but also logically. If one asks what order the principle of
sufficient reason introduces into nature, and
46 THE CAUSAL THEORY OF TIME

if one arranges phenomena in a series such that every term contains the
reason for all those which come after it in the series, the causal order of the
pheno-mena so defined will coincide with their temporal order of succession.
Now, if the temporal order of phenomena were not a primitive fact,
irreducible in the field of phenomena, but rather a simple consequence of a
more funda-mental order between the instants of absolute time which are
occupied by these phenomena, then the coincidence of the causal order of
phenomena with their temporal order would not yet permit an identification
of time with causality. But we have just seen that for Leibniz time is only the
order of phenomena; that is why he was able to propose the following
'metaphysical' defmition:

If one of two states which are not simultaneous involves a reason for the other, the former is held
to be prior, the latter posterior. 13

Here is how Leibniz deduced the genesis of the temporal order from his
defmition:

My earlier state involves a reason for the existence of my later state. And since my prior state,
by reason of the connection between all things, involves the prior state of other things as well, it
also involves a reason for the later state of these other things and is thus prior to them.
Therefore whatever exists is either simultaneous with other existences or prior or
posterior. 14

This extremely remarkable passage contains in nuce the whole causal theory of
time. In order to demonstrate that between any two instantaneous events
whatever there is always one of the three mutually exclusive relations of
posteriority, priority, and simultaneity, the classical substantialist [anti-relational]
theory had recourse to a linear image: the points on a straight line are to the right
or left of a given point, or they coincide with it. It is the same with instants
oftime, provided that the terms 'to the left of,' 'to the right of,' and 'coincidence'
are replaced with priority, posteriority, and simultaneity respectively. Events
possess the linear order of the instants which they occupy. Thus the ordinal
properties of absolute time explain the nature of temporal relations between
events and, in particular, the fundamental fact that every event must be posterior,
prior, or simultaneous in relation to every other. Now Leibniz replaces the
substantiaHst explanation of this fundamental fact, which had been made illusory
by his relational theory, with the causal structure of becoming and, in particular,
with universal interaction. He clearly supposed that physical action propagates
itself instantaneously from one substance to another, since he admits the
existence of a connection between
LEIBNIZ AND THE BEGINNINGS 47

simultaneous events. But it is easy to see that all that is required for the valid-ity
of his conclusion is the existence of physical processes which propagate with a
finite velocity as great as one may wish. Let us represent the successive states of
two substances by two constantly ascending curves and agree that two
simultaneous events are always on the same horizontal; that an oblique line
segment, joining the two curves, symbolizes a physical action propagated from
the lower point to the higher point; and fmally that the horizontal and vertical
distances measure space and time respectively. The velocity of pro-pagation will
be measured by the cotangent of the angle formed by the oblique and horizontal
line segments. In order for each point of the first curve to be joined to each point
of the second (which is not on the same horizontal) by an oblique line, it is
necessary and sufficient that there be line segments of as slight an inclination as
one may wish, i.e., that the velocity of propagation be as great as one may wish.
(See Fig. 1.)

cause

Fig. I.

In this way the linearity of the temporal order is found to be equivalent to a


property of causal action, namely that of the existence of actions propagated
instantaneously, or, at the very least, that of the non-existence of a maximum
velocity of causal propagation. It is clear that this linearity of the temporal
48 THE CAUSAL THEORY OF TIME

order, deduced by Leibniz from a hypothesis concerning causal action, could


be questioned, since his hypothesis (or rather the more limited one which we
have just substituted for it) has been refuted by the existence of a maximum
speed of propagation - which obliged science to replace the linear temporal
order with a 'conical' order (cf. Chapter IV, Section 1 below).
Leibniz also provided a 'metaphysical' defmition of simultaneity:

If a plurality of states of things is assumed to exist which involve no opposition to


each other, they are said to exist simultaneously. Thus we deny that what occurred last
1S
year and this year are simultaneous, for they involve incompatible states of the same thing.
This seems clear enough when one thinks of the states of the same substance
which cannot be simultaneous if they exclude one another; thus the states of
age and youth can never coexist in the same living being. But how can state
E1 of substance Sl exclude state E2 of substance S2? According to the classical
formulation, contradiction which is the coexistence of contradictory properties in
the same substance clearly assumes the identity of the substance (it is so,
moreover, in the above-mentioned example of the non-simultaneity of the
seasons). Perhaps the answer to this question should be sought in Leibniz's ideas
about logical attribution. For Leibniz, of course, every attri-bute is immanent in
the substance while on the other hand, all judgments concerning substances are
fundamentally attributive. 16 It is the same for relational judgment, which must be
decomposable into two attributive judgments concerning the terms of the relation
in question. Two relational judgments will be incompatible if the two pairs of
attributive judgments into which they are analyzed are excluded from the interior
of the two substances which form the terms of the relations. Thus, to cite another
of Leibniz's examples: if Mr. X is staying in India and his wife dies in Europe, his
widow-hood and her death will be simultaneous, or, if you like, the life of his
wife and his widowhood will not be simultaneous, because the judgment
concerning the life of Mr. X's wife can be analyzed into two attributive
judgments, one of which will attribute to Mr. X a property which is incompatible
stricto sensu with his being a widower.

Thus the Leibnizian theory of simultaneity, which can also be found in


slightly changed form in other representatives of the causal theory of time
(Kant, Lechalas, and Robb), assumes the classical formulation (incorrect in
our opinion) of the principle of contradiction, which here surreptitiously
introduces the notion of time. We shall return later to this temporal appli-
cation of an extra-temporal principle. It should be noted that in any case
Leibniz's treatment of simultaneity is based on a plan other than succession:
LEIBNIZ AND THE BEGINNINGS 49

in deducing succession from the notion of sufficient reason and simultaneity


from the notion of contradiction, he has done nothing but give a dynamic causal
explanation of succession. We shall soon see that Kant's theory, which provides a
homogeneous causal explanation of both succession and simulta-neity, makes
important progress over Leibniz's in this respect. (Furthermore, Leibniz's
affirmation of the instantaneous and universal connection implicitly contains the
Kantian definition of simultaneity by interaction.) Another important advance
made by Kantian theory is its epistemological spirit, in contrast to the
metaphysical spirit of Leibniz's. This metaphysical spirit is apparent not only in
his theory of simultaneity, which rests, as we have just seen, on the hypothesis
that all judgments are essentially attributive (which is nothing but a logical
translation of the metaphysical principle that the monads do not communicate
with each other: they 'have no windows'), but it is also apparent in his theory of
succession: the universal connection of things at a given instant, used in the
deduction of the temporal order, is again only the translation of the metaphysical
principle that each monad reflects the instantaneous state of the universe (Le., the
set of other monads) at each moment. This metaphysical spirit, so characteristic
of the whole Leibnizian philosophy of nature, is not surprising, especially in his
ideas on the 'meta-physical origins' of time. But it makes an epistemological
appreciation of Leibniz's theory difficult - what time is he talking about? On
what, in the last analysis, is his causal theory based? What is its epistemological
signi-ficance? We might be inclined to answer the first question by invoking
Leibnizian idealism: time is a causal order of phenomena, so time itself has only a
phenomenal import - which may give rise to the difficulty 17 of an exact
interpretation of this idealism. We might perhaps answer the second question by
recalling Leibniz's rationalism: our knowledge of the causal nature of time is a
priori knowledge; it proceeds by an analysis of complex concepts into their
simpler components - which does not take us far, given the unverifIable
metaphysical hypothesis on which the whole theory rests. Finally, as to the third
question, there is no way of telling whether we must interpret Leibniz's theory as
a rationalization of our ideas on the temporal structure of reality, or rather as a
'not inelegant' (as he liked to say) technical procedure which allows us to reduce
the number of indefmable 'mathematical' terms to a minimum.

Leibniz is then the precursor rather than the creator of a full causal theory of
time. This theory, which is essentially relational,18 (since it deduces temporal
relations between events from their causal connection without concerning itself
with the order of instants occupied by them) found the
50 THE CAUSAL THEORY OF TIME

ground cleared by Leibniz's relational theory, which showed that to attempt


to investigate the nature of time is to analyse temporal relations. Leibniz
defined an important temporal relation, succession, in terms of sufficient
reason, and indicated how this definition was linked to the causal structure of
the universe. But he did not extend his causal explanation to other temporal
relations, he neglected the epistemological aspect of the problem, and pro-
bably did not realize the essential novelty of his theory. We shall see how
Kant's theory, a result of quite different considerations, filled in important
lacunae in Leibniz's theory.
However, let us note, in Leibniz's case, an important feature which will
persist throughout the further development of the causal theory of time,
namely that the essential properties of the temporal and causal order, what-
ever their metaphysical origin (for Leibniz) or their physical origin (for his
successors) may be, are accepted as such by this theory. What it aims to do is
not to add new properties, but to clarify the nature of our knowledge of time
and causality by showing what are the fundamental elements of this
knowledge and what elements, on the other hand can be deduced from them.
The essential assertion of the causal theory - not only that temporal notions
are defmable in terms of the causal relations, but also, and more generally,
that all the properties of time can be deduced from certain properties of
causality - is implicitly in Leibniz. Its complete proof (that is, the construc-
tion of an axiomatic system whose axioms would be concerned only with
causality, while its theorems and defmitions would express the essential
properties of time), found in Leibniz in the form of a preliminary sketch, has
only been fully attempted in our own time. This proof is related to
epistemology, if by epistemology we mean all problems of acquired knowl-
edge, whether metaphysical or scientific. Leibniz's theory can very well be
characterized as epistemological (more specifically as axiomatic) as we
stated in the Introduction. Nevertheless, it does leave out problems on which
Kant would work, which are, strictly speaking epistemological (such as the
problem of the phenomenal nature of time) while post-Kantian philosophers
would develop Leibniz's axiomatic ideas instead.
II

KANT'S PHENOMENALIST INTERPRETATION OF


THE CAUSAL THEORY OF TIME

1. GENERAL REMARKS ON KANT'S PHENOMENALISM

We usually associate Kant's illustrious name with his idealist theory of space
and time, as expounded in the Transcendental Aesthetic. Indeed the first and
most impressive part of this masterpiece is explicitly devoted to the space-
time problem. But the Aesthetic contains only the synthetic a priori aspects of
the theory of sensory knowledge. Pure understanding, another source of our
knowledge of the phenomenal world, must form the ordered objects of
scientific experience from the chaotic diversity extended in space and time.
The union of the forms of sensibility and understanding in the Transcendental
Analytic could shed new light. We shall see, in fact, that a new aspect of the
problem of time, identical in certain respects to that in the definitive theory of
Leibniz, appears in the Kantian interpretation of the principles of pure
understanding, and, in particular, in his 'analogies of experience', which deal
with first, the indestructibility of substance, second, the principle of causality,
and third, the principle of universal interaction. It is known that Kant deduced
the validity of these three principles from the fact that they contain a priori
conditions of all empirical knowledge; the transcendental proof of their valid-
ity consists in the fact that, being presupposed in each experience, they are
logically prior to experience. In the detail of the proof of the three analogies,
Kant has recourse to empirical knowledge of time: permanence, causality,
and the interaction of substances have an a priori basis because, without them,
empirical knowledge of time would be impOSSible. We can understand
therefore why, in these pages of the Analytic, Kant was able to round out his
interpretation of time as a form of sensibility in a very important way. In
what follows we shall deal only with the second and third analogies, where
Kant expounds his causal theory of simultaneity and succession. The first
analogy does indeed contain a theory of the 'third mode of time', i.e.,
duration, but without a causal explanation.
The order of ideas expressed in the last two analogies is quite confused in the
text of the Critique, where Kant returns many times to the same subject, rather
capriciously intersperSing arguments, criticisms, and answers to possible
objections. We shall analyze these Kantian ideas from the point of view of the
51
52 THE CAUSAL THEORY OF TIME

causal theory of time, deliberately neglecting their primary aim in the author's
mind, which was the proof of the principles of causality and universal inter-
action. And, most important, we shall try to translate Kantian ideas into modem
language, knowing full well that the aphorism, traduttore-traditore, is especially
applicable in this case. Kant's language, as we all know, is semi-realistic, semi-
idealistic and systematically ambiguous, plunging the reader into a sort of
theoretical fog where it seems possible to counter each passage favorable to a
certain interpretation with another passage clearly unfavorable to it - precisely
what is responsible for the abundance of realist, idealist, and phenomenalist
philosophies, each of which, for different interpreters, bears the name of
Kantianism. It stands to reason that by using a more precise modem terminology
with clearer defmitions, we will artificially dissipate the Kantian fog, and that
this very clarification of Kant's ideas may change their character. However, in
this study we propose a critical analysis, rather than a historical reconstruction, of
the doctrines: it is impossible - or, if one prefers, too easy - to criticize a badly
defmed doctrine. It is up to us, then, to choose from among the different
meanings attributable to Kant's text that which, to all appearances, most
conforms to the general spirit of his doctrine, and lends itself least to criticism. It
is with this one reservation that we present the reader with the following
exposition of the Kantian theory.
The distinction between subjective and objective time, familar to every reader
of Newton, is Kant's starting point. In his theory, where time has no meaning
outside of human representation, this important and inevitable distinction should
be interpreted in terms other than those of contemporary realism, where objective
time is the time of things in themselves. We are aware, he says, that our
perceptions succeed one another in time: how do we come to know that the
corresponding objective states also succeed one another in the object? If, in
looking at a house, I let my eye wander from top to bottom, from right to left, the
perceptions of the different parts of the house succeed one another - but
obviously I should not conclude from this that the parts of the house succeed one
another, since they are all simul-taneous. Consequently, succession in perception
does not always involve succession in the perceived reality. The same will not be
true of the perception of a boat sailing down a river: to the successive perception
of the different positions of the boat there will correspond, 'objectively', a
succession of these same positions; in this last case, the subjective temporal order
of perceptions is paralleled by an objective order of perceived events, while in
the first case the two orders were dissociated. Why this difference?

In order to understand the answer supplied in the Critique, it is especially


KANT'S PHENOMENALIST INTERPRETATION 53

important to grasp the sense of this Kantian question. Since we are given
only our representations, all empirical knowledge must involve the
representations, present or future, real or virtual, but never leave the realm of
possible experi-ence. Therefore, if the distinction between subjective and
objective time is based on experience, it must be interpreted as a function of
representations. If the succession of perceptions of the positions of the boat has
an 'objective' significance completely different from that of the perceptions of the
different parts of the house, it is because the first group of representations has
pro-perties totally different from those of the second. Now, in a fundamental
passage,19 Kant asks what is this new property acquired by representations (of
succession, in particular) as they become objective, and conform to an object? It
cannot be the relation to another presentation, since the question would then arise
for this one as well. The only consequence of the objectivity of
representations is to render necessary the existence of some connection
among them; and vice versa, the only means we have to assure an
objective value for our representations of succession is to establish some
necessary connection among them.
For example, the difference, for Kant, between yesterday and a dream
does not lie in the fact that real objects correspond only to yesterday's repre-
sentations, but rather in the fact that the connection between yesterday's
representations is different from that of the dream. This criticism does not
deny that this immanent difference may be accompanied by a metaphysical
difference with a bearing on things-in-themselves: what it affIrms is that for
human knowledge the connection of representations is the only means of
establishing their objectivity, that it is also sufficient.
This profound thought of Kant's may be illustrated by an example taken
from the theory of cardinal numbers, made famous by the work of Bertrand
Russell. The statement that two arbitrary sets have the same number of
elements can be interpreted in two ways, because the equality of the sets,
which is generally demonstrated by the existence of a cardinal number
common to both and distinct from each, could also be established by a direct
comparison of the two sets (establishing a unique and reciprocal one-one
relationship between their elements) without considering the existing of a
common number. We can see the epistemological advantage of the second
method: in making the hypothesis of cardinal nwnbers superfluous as distinct
from the sets of which they measure the power, it alone satisfies Occam's
postulate. If we replace the sets with our representations, their equality with
some 'objective' phenomenon, and the cardinal number with the thing-in-
itself, we can understand how, without dogmatically denying the existence of
54 THE CAUSAL THEORY OF TIME

the thing-in-itself, Kant wished to explain the objectivity of phenomena by


means of special 'correspondences' (in this case not one-one but 'necessary')
among the representations.

2. THE CAUSAL THEORY OF SIMULTANEITY

These relations differ in fact in the two types of perceptual series. One series is
nothing but a subjective succession of psychic states, while the other corre-
sponds to an objective succession: the order of the perceptions of the different
parts of the house is arbitrary, while that of the perceptions of the moving boat is
necessary. True, my eyes scanned the house in a certain determined order, say
from top to bottom, or from right to left; but I could just as well have surveyed
the house in the opposite order without the perceptions themselves being
changed. The same is not true of the second example: I could not have perceived
the different positions of the boat in an order different from that which actually
occurred; the perceptions here follow one another in a determined order. Thus,
according to Kant, objective succession and objective simultaneity are both
represented by subjective successions. (He does not seem to have analyzed the
epistemological Significance of subjective simultaneity - for example, the
intuitively perceived simultaneity of two sounds - which is certainly an
irreducible and fundamental intuitively given fact.) Nevertheless, we must not
conclude from this that, for Kant, ''to perceive an objective succession" would
only be a convenient manner of speaking, whose real meaning would be ''to pass
through an irreversible series of psychic states" (such as the series of perceptions
of the positions of the boat) and that "to perceive an objective simultaneity" really
means "to pass through a reversible series .of states" (e.g. the series of
perceptions of the house). This interpretation, suggested by some passages of
Kant's text and partially adopted by Schopenhauer, does not take sufficient notice
of the role played by the rationalist factor in Kant's phenomenalism. The
following passage dealing with simultaneity seems to us to be decisive:

Things are coexistent when in empirical intuition the perceptions of them can follow upon
one another reciprocally, which, as has been shown in the proof of the second principle,
cannot occur in the succession of appearances. Thus I can direct my perception first to the
moon and then to the earth, or, conversely, first to the earth and then to the moon; and
because the perceptions of these objects can follow each other reciprocally, I say that they
are coexistent. Now coexistence is the existence of the manifold in one and the same time.
But time itself cannot be perceived, and we are not, therefore, in a position to gather,
simply from things being set in the same time, that their perceptions
KANT'S PHENOMENALIST INTERPRETATION 55

can follow each other reciprocally. The synthesis of imagination in apprehension would
only reveal that the one perception is in the subject when the other is not there, and vice
versa, but not that the objects are coexistent, that is, that if the one exists the other exists at
the. same time, and that it is only because they thus coexist that the perceptions are able to
follow one another reciprocally.2o Consequently, in the case of things which coexist externally
to one another, a pure concept of the reciprocal sequence of their determinations is required, if
we are to be able to say that the reciprocal sequence of the perceptions is grounded in the object,
21
and so to represent the coexistence as objective.

Kant's thought is clearly the following: the fact of having successively per-ceived
the earth and the moon in an arbitrary order does not immediately asssure us of
their simultaneity, since it is always possible that when one of them enters the
field of consciousness the other is destroyed, and vice versa, so that they are
never simultaneous, although the order of their perceptions is arbitrary. Therefore
a new element must be added to this type of perceptual series in order for them to
be considered as 'perceptions' of an objective simultaneity (in fact, for Kant there
is, strictly speaking, no perception of objective temporal relations; to say of a
perception that it is of an objective temporal relation always implies some
function of the understanding, which alone can objectify the representations of
sensibility). This new element is the concept of interaction, furnished by the
understanding: for two objects to be simultaneous, there must be interaction
between them. Nevertheless, as we said above, for Kant every judgment about the
empirical object is resolved into judgments about certain representations
(precisely those which common sense, with its 'naive realism', would call 'the
representations of the object' if it were to become aware of the existence of
perceptual representations). The same must be true of the interaction between
objects, which must be completely expressed by a rule for the synthesis of
representations. In the case under consideration it is the rule that at each instant I
could have observed the earth, as well as the moon, so that the series of my
perceptions, although perfectly determined, would be virtually replaceable with a
completely different series of perceptions of these objects, as proven by the
perceptions of other observers.

Clearly, the fact that a perceptual sequence is virtually replaceable by another


is no longer a given of sensibility. To verify this we must have recourse to the
concept of interaction. Kant did not explain in detail how the reversi-bility of a
perceptual sequence is related to objective interaction. Perhaps he would have
given the following explanation: if substance S1 acts on substance S2 during an
interval of time, and if I perceive them both in turn, I am certain that substance S
1 existed even when I perceived only S2, since the state of
56 THE CAUSAL THEORY OF TIME

8 2 , of whose existence I was assured by my perception, because of


interaction, entails the existence of the corresponding state in 8 1 which,
consequently, could also have been perceived. In asserting the interaction of
two substances, I then deduce that the actual order of my perceptions could
have been reversed at each instant, and it is in this reversibility of my
representations that I recognize the simultaneity of the objects. But we must
not conclude that the reversibility of a perceptual sequence is sensibly given
in the same sense as is every perception of the sequence. Only the
perceptions are given; the relation of interaction is asserted by the
understanding so as to link the objects of perception.
This interpretation, which is not rigorously idealistic, since it admits the
validity of empirical laws as a given, irreducible to representations although
immanent with regard to them, seems to us to take better account of the
rationalist component of Kantianism. Of course, the third analogy does not
deal with universal interaction conceived as an indefinable category of pure
understanding, but only with its applicability to the realm of phenomena,
where, in conformity with the 'schematism of the pure understanding', the
universal interaction is reduced to a rule of synthesis of representations: but
the schematism 'restrains' and 'realizes' the categories without defining them.
Moreover, in affIrming that the categories of the understanding, and in
particular that of interaction, are applicable only to phenomena, Kant does not
maintain at all that they are applicable to the representations themselves: he
distinguishes clearly between the phenomena and representations, the former
being a rule applied to the latter which, thus, follow the rule. Today we would
say, speaking from the point of view of the extension of the concepts being
analyzed, that, according to Kant, the phenomenon is quite simply the set of
representations (real or virtual) which follow a given rule. This explanation of
the phenomenon is undoubtedly inadequate, since there are rules of
representations which are not phenomena; the thing is not just any set of
representations, even for the phenomenalist. In order to complete Kant's
explanation, we would need special research of the sort which led Russell to
defme the thing as the set of its aspects, and Husserl to his in-teresting views on
the 'constitution of the thing'. But Kant's general tendency seems suffIciently
clear to us, and when applied to our particular case, it may be formulated thus:
subjective time is an order of representations, objective time is an order of
phenomena; subjective simultaneity is not defmed by Kant, objective
simultaneity is defmed in terms of interaction. This interaction among objects is
translatable into a property of their representations: they form a reversible
sequence. But the reversibility of a perceptual series is no
KANT'S PHENOMENALIST INTERPRETATION 57

longer a sensible given: it is only the equivalent of a dynamic relation among


phenomena, posited by the understanding. For the order of two perceptions
to be reversible, it is necessary and sufficient that there be interaction be-
tween their objects.

3. THE CAUSAL THEORY OF SUCCESSION

This point is brought out still more clearly in the Kantian theory of suc-
cession. In order for two phenomena to be called successive, the temporal
order of their perceptions must be necessary, i.e., the inverse order must be
impossible. It is impossible to perceive the downstream position of the boat
before its upstream position, and this is why the latter is objectively prior to
the former. Objective succession is not to be confused with the succession of
perceptions; the new added element, necessity, or its counterpart, impossi-
bility, is brought into' play only as a function of understanding: sensibility
and 'imagination', in the Kantian sense, can ascertain that perception A was
prior to perception B, but they cannot assure us that this was necessary, that
its converse was impossible. Only the causal relation can verify this fact.
Using Kant's example, if I put a lead ball on a cushion, the depression in the
cushion will be determined by the smooth shape of the ball; however, if the
cushion already has a depression in it, received I don't know when, I would
not be able to conclude from that the existence of a lead ball. 22 Thus, it is by
asserting a causal relation that I can be assured that the perception of the
depreSSion necessarily (and consequently, objectively) followed the
placing of the ball.
Now we can see how Kant inverted Hume's problem: for Hume the causal
relation is only a regular succession. Knowledge of it is purely empirical: the
mind establishes that two phenomena have followed each other to date, and
concludes that they will always do so. But what sort of succession is he
considering? Clearly an objective succession. According to Kant, objective
successions presuppose causality and cannot serve to define it. In Hume's
view, to establish that a phenomenon A is the cause of a phenomenon B, one
must repeatedly establish that B follows A. But, according to Kant, in
establishing that B has objectively followed A, we thereby assert a causal
relation. It was therefore a new solution to Hume's problem that furnished
Kant with his causal defInition of succession.
In other respects, the analogy between succession and simultaneity seems
perfect: just as the interaction of two objects is equivalent to a reversible se-
quence of perceptions of them, causality amounts to an irreversible sequence,
58 THE CAUSAL THEORY OF TIME

i.e., a necessary succession of perceptions. To say that an objective


succession corresponds to a necessary subjective succession, subject to a
flxed rule, is to say that causality and objective succession are translatable
into the sphere of representations by the same type of synthesis. But, for a
phenomenalist, this amounts to confusing causality with objective
succession, i.e., to adopting the Leibnizian causal theory of succession and
extending it to the 'second mode of time', simultaneity. But Kant himself did
not explicitly draw this conclusion from his insights into objective succession
and simultaneity in order to identify objective time with causal order. Yet the
conclusion follows immediately and inevitably from any attentive analysis.
Schopenhauer, who made a remarkable criticism of the Kantian theory -
which we will soon consider - formulated the conclusion thus:

If objectivity of succession were known merely from causality, it would be conceivable only as
such, and would be nothing but this ... following and being effect would be one and the same
23
thing.

Let us note, in addition, that in Kant the causal deflnition of simultaneity is


much less precise than that of succession. At first sight the necessary and
suffIcient condition for two objects being simultaneous is the existence of a
reciprocal action between them. But here is a passage which seems to
suggest a less paradoxical interpretation:

We may easily recognize from our experiences that only the continuous influences in all
parts of space can lead our sense from one object to another. The light, which plays
between our eye and the celestial bodies, produces a mediate community between us and
them, and thereby shows us that they coexist. We cannot empirically change our position,
and perceive the change, unless matter ill all parts of space makes perception of our
position possible to us. For only thus by means of their reciprocal influence can the parts
of matter establish their simultaneous existence, and thereby, though only mediate-ly, their
coexistence, even to the most remote object. 24

This seems to suggest that two objects which do not act directly on each
other can nevertheless be considered simultaneous, on the condition that both
act jointly on a third object. We will return to these different possibilities of
deflnition when we come to study some recent theories, whose authors took
the care to formulate their definitions with a precision capable of satis-fying
the requirements of a rigorous logic. Obviously, this is not true of Kant's
exposition, and it would be wasted effort to examine his propositions under a
logistic microscope. The essential point here is that, for Kant, simultaneity is
a vaguely deflned combination of causal relations.
KANT'S PHENOMENALIST INTERPRETATION 59

4. EXAMINATION OF SCHOPENHAUER'S CRITICISM

Kant's ideas on the causal theory of time have often been the target of more
or less justified criticism.25 We shall analyze briefly the most comprehensive
of these, expounded by Schopenhauer in a well-known section of his doctoral
dissertation (,23. Bestreitung des von Kant aufgestellten Beweises der
Aprioritiit des Kausalgesetzes' ['An objection to Kant's proof of the a priori
aspect of causal law']. Schopenhauer's objections are directed almost exclusively
at the causal definition of succession; he rejects at the outset the definition of
simultaneity in terms of causal interaction, considering this a contradictory
concept - mistakenly, in our opinion, since in Newtonian physics, with which
Kant agreed, the law of universal gravitation as instantaneous propagation firmly
established a universal interaction in the Kantian sense. The following are
Schopenhauer's main objections to the causal nature of succession:
(a) The Kantian examples of the perceptions of a house and a moving
boat, according to Schopenhauer, both concern objective successions. The
temporal order in which we perceive the different parts ofthehouse (although
they are simultaneous) is every bit as objective as that in which we perceive the
successive positions ofthe boat going down the river: each is the result of a
relative motion, the first, of the motion of the eye with respect to the house, and
the second, of the motion of the boat with respect to the observer; the voluntary
character of the first has no importance with regard to time. But Schopenhauer's
objection is not to the point, since in neither example does Kant deny the
succession of perceptions as subjective states: what he does assert is that in the
first example an objective simultaneity, that of the co-existent parts of the house,
corresponds to a subjective succession, while in the second, the subjective
succession is paralleled by an objective one. The
Kantian example therefore proves conclusively that the two perceptual series
do not have the same objective significance, and that from the point of view
of the phenomenalist this difference must be grounded in phenomenal dif-
ferences between the two series. But it seems doubtful that it is a difference
between simultaneity and objective succession. Undoubtedly, the parts of the
house coexist, while the positions of the boat follow one another. But it
doesn't follow from this that the parts ofthe house which I successively per-
ceived also coexist. This inference obviously presupposes the 'numerical'
identity of the object, perceived twice in succession. Yet it seems more in
conformity with the present state of our knowledge not to consider the
duration of a substance as implying its numerical identity through this dura-
tion. We will analyze this point in detail in connection with Russell's theory.
60 THE CAUSAL THEORY OF TIME

For the moment, let us simply note that a suitable modification of Schopen-hauer's
criticism brings to light an important gap in Kant's theory, which provides a causal
explanation of succession and simultaneity, but not of duration. It also appears
doubtful that the simple property of reversibility suggested by Kant is enough to
distinguish the two types of perceptual series: for example, the objective simultaneity
of two instantaneous events cannot be established by the Kantian method, since,
lasting only an instant, it is impossible to perceive them one after the other - unless
one brings into play the time taken for the propagation of light or any other physical
process which serves as intermediary between the event perceived and the body of the
observer. Besides, this is a delicate point, since Kant, as Schopenhauer points out,
does not seem to have taken propagation-time into account. 26 Clearly the existence of
a fmite velocity of propagation very much complicates the interpretation of the
perceptual process. Still, suitably modified, the Kantian theory could be adapted to
this complicated state of affairs. If the speed of light were infmite, the correspondence
between subjective and objective time in the Kantian theory could be represented by
the two diagrams in Figure 2.

s imultanei ty succession

Fig. 2.

In these diagrams the 'R's designate the representations, the 'O's corre-
sponding objects (we just saw how Kant explains this correspondence in a
phenomenalistic sense), the vertical arrows symbolize the relation of corre-
spondence between representation and object, and fmally, the horizontal
arrows symbolize the subjective temporal order of the representations. In the
frrst diagram, objective simultaneity corresponds to a reversible sequence of
representations (subjectively capable of taking place in both directions); in
the second, the irreversible subjective succession defines an objective
succes-sion. Such is Kant's theory, based on instantaneous propagation.
KANT'S PHENOMENALIST INTERPRETATION 61

If we reject this, things become complicated, since the objective temporal


order of two events can be the converse of that of their representations, provided
the events take place in suitably chosen locations; the reversibility of the order of
representations no longer guarantees the simultaneity of the corresponding
events. To maintain simultaneity, it would be necessary to interpret the length of
the arrows as a measure of the time interval and, supposing subjectively that Rl
preceded R 2 , to make R 2 0 2 as long as the sum of R 1 R2 and RIO 1. But this
clearly presupposed knowledge of the speed of propagation, or, in the final
analysis, knowledge of objective time. There seems, then, to be a vicious circle in
any determination of an objective temporal relation, sirlce it presupposes the
measurement of a velocity, and this irl tum presupposes some objective temporal
facts. We know how this dilemma is resolved in practice: we act as if the
propagation were approxi-mately irlstantaneous for relatively short distances.
This supposition cannot be proven by further measurement, since it is
presupposed irl all determirlations of objective time: if I do not assume that the
telescope, which I n0W see and by the aid of which I will be able to calculate the
spatio-temporallocation of some stellar event, was irl such and such a place a
while ago, I will never prove it. This is a case of an a priori hypothesis (or rather
a typical convention) which, under one form or another, is irlevitable irl any
cognition of what is real. Given this assumption, we accept irlstantaneous
propagation withirl a certain domairl, and there the Kantian criteria become
applicable. To go beyond this irrltial domairl, it will again be necessary to use the
causal relations with the rest of the universe, i.e., those irlfluences and
irlteractions which lirlk the diversity perceived by the subject at a given irlstant to
the most distant regions of space and time. It is along these influences and
irlteractions that the temporal structure of what is perceived is transmitted, until
fmally it embraces the whole universe. Along the way this structure will lose this
semi-intuitive character, with which it was endowed within the irrltial domairl by
the series of perceptions on which it was based. What, in this domairl, was
objective simultaneity paralleled by a reversible perceptual sequence becomes,
outside of it, just irlteraction; what was objective succession, paralleled by an
irreversible subjective succession, becomes asymmetrical causality. Thus, there
would never be objective simultaneity without irlteraction, nor objective
succession without causality, since within the irlitial dornairl, as well as out-side
of it, the objective temporal order would be derived from certairl causal
considerations (cf. Chapter VI, Section 3).

(b) Schopenhauer next objects that there are many objective successions
which do not have a causal character: for example, the sounds of a melody
62 THE CAUSAL THEORY OF TIME

follow one another without the earlier one being the cause of the later one;
or, if I leave a house, and a brick, sliding off the roof, falls on my head, it will
not be my leaving that caused the fall, although it was prior to it (since all the
evidence points to the conclusion that the brick would have slid, even if no
one had left the house). It is therefore possible to establish that a succession
took place without appealing to causality. - What could a partisan of Kant
reply? In the first example he would distinguish the succession of auditory
sensations from the succession of sounds, considered as physical phenomena.
The former, the only one which is immediately perceivable, has to do with
subjective time and is not covered by the causal theory; the latter can cer-
tainly not be established without recourse to causal considerations. In the
second example, the Kantian would undoubtedly ask fIrst how this
succession without causality could have been known. The brick, (as) physical
object, is a certain group of representations entirely subject to a common
objectifying rule; the same is true of the body of the man leaving the house,
considering him only as a cause in this case. If someone saw the man leaving
before the fall of the brick, it means that a certain representation of the group
'hwnan body' preceded (subjectively) another representation of the group 'brick'.
This succession must have been irreversible for the observer to have been able to
conclude that it was an objective succession. But, in this case, there was a causal
relation between the two objects. (This might be the Kantian answer to
Schopenhauer's objection.) It is unacceptable, disappointing, for Schopen-hauer,
who, though in principle an idealist, nevertheless conceives of physical objects as
a realist, and is not aware that for Kant they are nothing but rules for the synthesis
of representations. It is true that Kantian phenomenalism is not a completed
doctrine, but rather a magnillcent program whose realization could bring fame to
a whole generation of scholars. This program will be completed when our whole
science of the real will be expressible in terms of immediate data and rules of
objectifying synthesis. Similarly, it seems to us that the causal theory of time, in
the phenomenalistic sense, represents nothing but a research program, realizable
on the basis of a complete phe-nomenalistic construction, of which we have only
the beginnings here. Kant's followers will try above all to eliminate his 'virtual'
representations which are hardly representations at all and which contradict the
very principle of phenomenalism. In constructing the causal theory of time they
will use the Kantian indications to mean that the objectifying rules dealing with
the temporal order are of the same kind as those dealing with the causal relation;
it is evident that the causal connection, insofar as it is 'objective', must be
represented in the domain of representations by suitable rules.
KANT'S PHENOMENALIST INTERPRETATION 63

Going still further, the answer to Schopenhauer's objection can be put into a
form independent of Kant's phenomenalism although still in conformity with his
causai theory of time. In fact it is easy to show that the answer emerges from the
principles of causality and universal interaction admitted by Kant. The fIrst
principle must be formulated in Kant's substantialist terms: the state of each
substance acts on every subsequent state of every other substance. And the
second: there is interaction between the two simultaneous states of any two
substances. (Let us note that, thus formulated, these two principles become the
defmitions of succession and simultaneity, respectively, while Kant considered
them to be synthetic a priori judgments. We will not examine this disagreement,
unnoticed by Kant and perhaps only apparent.)
Having said this, let us return to Schopenhauer's example. My leaving the
house is a state of the substance called my body; the fall of the brick is a
posterior state of another substance. According to both the principles of
causality and of universal interaction, there must be an influence emanating
from the fIrst state and reaching the second. It is precisely by following this
influence coming from the frrst state and ending in the second that, according
to Kant's hypothesis, I can infer a relation of objective succession between the
two substantial states. For example, in the particular case of the perceived
succession where light, 'playing' between the object and the observer's body,
according to Kant, creates, a dynamic interaction between the successively
perceived events El and E2 and the states E1 ' and E2' of the observer's body,
which are respectively simultaneous with them, the causal chain E 1 , El ',E2 ',
E2 links anterior event El to posterior event E2 through the body of the
observer. It appears then that in Schopenhauer's example, there is indeed a causal
action by their earlier event upon the later one, whether the succession be
perceived or inferred. What gives Schopenhauer's argument its apparent force,
and also Kant's defmition its apparent strangeness, is the fact that Kant's
defmition seems to presupposes that every event contains the complete cause of
every subsequent event. But it is clear that this is not Kant's opinion; he is only
thinking of an influence, i.e., a partial determination. Basically, for Kant,
succession is equivalent to a partial asymmetrical determination, simultaneity to a
partial symmetrical determination. In our particular ex-ample, this means that in
order for the man's leaving to be prior to the fall of the brick, it is necessary and
sufftcient that a more extended event of which the leaving forms a part be the
cause of an extended event of which the fall is a part. Thus the diffIculty, noted by
Schopenhauer, vanishes, a difftculty which would have attributed to Kant an
opinion which he did not hold at all.
64 THE CAUSAL THEORY OF TIME

(c) Schopenhauer's third objection is that if Kant's theory were true we


could not know the reality of succession without knowing its necessity,
which could be possible only for an understanding which perceived all the
causal series at once, in other words, an omniscient understanding. "Kant has
burdened the understanding with an impossibility merely in order to have less need
of sensibility." 27 This objection means in effect that we often think we know some
real successions with all possible certainty, while it seems im-possible to establish
their causal necessity, since in order to do so it would be necessary to know all the
successions of the same type. But let us recall once more that Kant is thinking of
knowledge of the objective universe: this knowledge is always inductive and always
goes beyond the immediate given. Knowledge of a simple objective succession in
principle raises the same difficulties as knowledge of a causal relation or of any other
objective fact, however complex. The causal definition of objective Simultaneity
makes it clear that it can only be inferred with probability, without us being able to
acquire immediate and unquestionable knowledge of it. This would in effect
necessitate the realization of an infinite number of perceptual sequences resulting
from an arbitrary inversion in a given sequence, which is impossible. According to
Kant knowledge of a succession or simultaneity rests on the sensible knowledge of a
subjective succession, but the two must not be confused with each other. Thus, it is
not correct to say, as Schopenhauer does, that Kant misunderstood the role of
sensibility in the knowledge of time. We will return to this problem when we discuss
Reichenbach's theory. For the moment, let us simply note the curious fact that the
reproach of misunderstanding the role of sensibility in the knowledge of time, made
by Schopenhauer to Kant, was also made by Kant himself to Leibniz:

If I attempt, by the mere understanding, to represent to myself outer relations of things, this can
only be done by means of a concept of their reciprocal action; and if I seek to connect two states
of one and the same thing, this can only be in the order of grounds and consequences.
Accordingly, Leibniz conceived space as a certain order in the com-munity of substances, and
time as the dynamical sequence of their states. That which space and time seem to possess as
proper to themselves, in independence of things, he ascribed to the confusion in their concepts,
which has led us to regard what is a mere form of dynamical relations as being a special
28
intuition, self-subsistent and antecedent to the things themselves.

This passage is especially remarkable for its summary of Leibniz's spatio-


temporal dynamism; it conflfms the fact that Leibniz's causal theory only
emphasizes a concept of time implicitly contained in his dynamism. But it
also reminds us of the two-fold change which Kant brought to Leibnizian
KANT'S PHENOMENALIST INTERPRETATION 65

dynamism: he interpreted it phenomenalistically and showed that it was


inadequate with respect to empirical knowledge, which is the synthesis of the
subjective time of sensibility and the dynamic time of understanding. (cf.
Chapter II, Section 5).

How can we reconcile Kant's assertion that the objectivity of succession is known only
from the necessity of the effect's following from its cause with his other assertion
(Critique ofPure Reason, 1st ed., p. 203; 5th ed., p. 249) that succession is the empirical
criterion as to which of two states is cause and which effect? Who does not see here the
most obvious circle? 29

This objection is of paramount importance for any causal theory oftime. If, in
order to distinguish cause from effect, it is necessary to know the temporal
relation between the two, it will never be possible to deduce the temporal order
from the causal order without a vicious circle. The passage where Kant asserts
that succession is the only empirical criterion for causality seems to us very
questionable. If Kant was thinking of objective succession, then Schopenhauer's
criticism is sound. But it could also be possible that Kant was thinking of
subjective succession; if that is the case, there is no vicious circle in considering
it as a criterion of causality, and causality as defining objective succession.
Everything depends, then; on how Kant conceived of the empirical knowledge of
a particular causal relation. To the best of my knowledge, nowhere in Kant is
there any clue to this. We will see subsequently how his successors, particularly
Lechalas and Reichenbach, answered this question.

5. THE PLURALITY OF CONCEPTS OF TIME IN KANT

Schopenhauer's four objections do not seem to be decisive with regard to Kant's


theory. Let us point out a fIfth objection which, although suggested by the text of
the Critique, apparently did not worry Schopenhauer:
How should we reconcile Kant's causal theory with his views expressed in
the Transcendental Aesthetic? If time is a pure intuition, how can it be identical
with the causal connection of phenomena? We have no intention of denying that
this presents a real diffIculty for anyone who attributes a causal theory of time to
the author of the Aesthetic. However, perhaps this diffIculty is not
insurmountable, and, without claiming absolute historical accuracy for what
follows, we shall briefly indicate how it might be removed; all that we claim is
that it is possible to coordinate the two Kantian concepts of time while still
taking into account his general theory of empirical knowl-edge. We know that in
his analysis of empirical knowledge, Kant clearly
66 THE CAUSAL THEORY OF TIME

distinguishes between intuition (Anschauung) and perception (Wahrnehmung);


man does have a pure intuition of time as a form of sensibility, but, strictly
speaking, he has no perception of it. This assertion that time cannot be perceived,
plays a decisive role in the proofs of the three analogies; its origin is easily
explained by the fact that Kant attributed to time the properties of Newton's
absolute duration, and this duration, "true, absolute, and mathe-matical", distinct
from "apparent, relative, and sensible" duration, is clearly not perceivable. Kant
adopts the causal theory to explain our empirical knowledge of absolute and non-
perceivable time. If time could be perceived like phenomena, the position of
phenomena in time would be immediately given to the senses and would not
require the aid of the understanding and the principles of causality and
interaction. Then there would be three distinctions to make: pure time, a form of
inner sense; subjective time, the order of representations; and objective time, the
order of phenomena. The fIrst is irreducibly given and immanent in the human
mind; it would certainly be contrary to the Kantian doctrine, which distinguishes
between understanding and sensibility as absolutely separate sources of
knowledge, to deduce time, a form of sensibility, from causality, a principle of the
understanding. There-fore, the causal theory is certainly not applicable to the
Kantian conception of pure time. Nor is it applicable to time considered as the
order of represen-tations, or more generally, of subjective states as such, because
Kant invokes causality to explain the transition from the subjective time of
representations to the objective time of phenomena; the fIrst is the point of
departure for and presupposed by the application of causality, and consequently
cannot be explained by it. Therefore, the causal theory is applicable exclusively to
objective time as distinct from pure time and the time inherent in the psychic flux.
This distinction has nowhere been explicitly made by Kant; nevertheless, it seems
to us in conformity with the general spirit of his doctrine and neces-sary to its
coherence. There are certain passages which could be interpreted as asserting the
identity of the pure time of sensibility with the objective time of phenomena: in
this case, the objective order of phenomena would be a consequence of their
respective situations in pure time, and causality would be necessary only for the
knowledge of this situation. But this distinction between 'being' and 'knowing'
does not seem to us to be in the spirit of the Critique, according to which
phenomena exist for us only insofar as they are known by us, and even, we might
add, as they are accessible by the methods of physics. In fact we should not forget
that the Critique ofPure Reason was throughout its positive portion 'an essay in
physical philosophy'. Its principal concern is with the possibility of synthetic a
priori judgements, which are in
KANT'S PHENOMENALIST INTERPRETATION 67

fact the principles of Newtonian physics. Newton's famous preliminary de-


fmitions, which made transcendent entities of space and time, are the key to
the whole Transcendental Aesthetic. However, as Kant's physical philosophy
is a critique of physical knowledge and not a metaphysics of the material
universe, absolute space and time., which were for Newton basically divine
attributes, became for Kant a form of human representation. The revolution
achieved by Kant, was less a Copernican (as he himselfliked to call it,) than
a Socratic one:

For time is not viewed as that wherein experience immediately determines position for
every existence. Such determination is impossible, inasmuch as absolute time is not an
object of perception with which appearances could be confronted. What determines for
each appearance its position in time is the rule of the understanding through which alone
the existence of appearances can acquire synthetic unity as regards relations of time; and
3o
that rule consequently determines the position.
It is indeed the rules of the understanding, subject to the principles of causality
and universal interaction, which engender the objective temporal order of
phenomena.
These are the physical, or more precisely, Newtonian, preoccupations of
Kant which appear in the threefold nature of time just expounded. What
seems to us to invite criticism in these ideas of Kant is not that they admit of
too many distinct times, but rather too few. To explain epistemologically the
difference between Newton's time and subjective time, Kant seems to have
had recourse to the causal character of the ftrst, as opposed to the sensible
character of the second. But is this suffIcient to understand every objectiftca-
tion of sensible time? The Kantian theory clearly presupposes that there is
only one objective time, namely that of physics. But how shall we classify the
temporal relations between different psychic fluxes? The simultaneity of two
physical events belongs to physical time, the simultaneity of my repre-
sentations, to subjective time. But to what time does the simultaneity of my
representation with that of my neighbor belong? Clearly, it can neither be
sensibly given, nor reduced to physical causality. It seems therefore that in
addition to subjective time and physical time, the only ones considered by
Kant, we must distinguish an intersubjective time in empirical knowledge,
and that if we neglect it we will never come to understand the historical
knowledge of intersubjective relations. Moreover, this knowledge orders
psychological phenomena with respect not only to one another, but also to
physical phenomena. This is what we ourselves do constantly when we con-
sider a movement as posterior to the decision to execute it, or the perceived
68 THE CAUSAL THEORY OF TIME

event as approximately simultaneous with our perception of it. This psycho-


physiological order, in constant use in psychology and physiology, entails
another objectification of subjective time, which does not fit into the Kantian
scheme. Thus, the temporal order has four distinct forms, depending upon
whether it relates two physical events; or two states of one consciousness; or two
states belonging to different psychic fluxes; or finally, two states, of which one is
physical and the other psychological. The Kantian theory, aiming at an
epistemological interpretation of Newtonian physics, makes a thorough study of
only the first form of this order, only touches upon the second, and completely
ignores the last two. The same is true of post-Kantian authors, who, because of
their interest in an epistemological inter-pretation of physical time, paid no
attention to intersubjective and psycho-physiological time and contented
themselves with a few summary remarks about subjective time, claiming that its
more thorough study belonged to psychology and not to epistemology -
mistakenly, in our opinion, since, mutatis mutandis, the same could be said
about physical time. We believe, however, that only a simultaneous and far-
reaching analysis of all the forms of temporal order, such as will be given in Part
II of this book, has any chance of making us understand the whole of our
empirical knowledge of time, and, in particular, the genesis of universal time,
which embraces every phenomenon, physical as well as psychological in a single
order. It may be that this will provide the explanation of the opposition between
the pure time of the Transcendental Aesthetic and the causal time of the Analytic.

Thus Kant's theory undoubtedly contains some important gaps. Never-theless


it made a decisive advance over Leibniz's causal theory. Leibniz treated only
succession in a causal manner while Kant, neglecting duration, gave a causal
explanation of succession and simultaneity, which agreed with the principles of
Newtonian physics. Leibniz neglected the epistemological problem; Kant
explained the epistemological genesis of the physical temporal order by
analyzing the role played by subjective time and causality in our knowledge of
physical time. No doubt even the data in Kant's problem will be changed when
science is forced to abandon Newtonian principles. Not only his causal
definitions, but also his phenomenalistic interpretation will then have to be
revised. It seems to us extremely regrettable that the subsequent development of
the causal theory of time has been almost exclusively re-stricted to the
epistemological component of Kant's theory (which was a continuation of
Leibniz's ideas), while the most durable part of Kant's work, namely his profound
although incomplete epistemological considerations
KANT'S PHENOMENALIST INTERPRETATION 69

on the relation between subjective time and physical time as well as the role of
causality in the knowledge (and not only the structure) of physical time, was not
developed by his successors.
III

LECHALAS' ADAPTATION OF THE CAUSAL


THEORY OF TIME TO THE LAWS OF PRE-
EINSTEINIAN PHYSICS

1. GENERAL REMARKS ON LECHALAS' THEORY

We shall not spend time on the study of the various post-Kantian authors, such
as Balmes 31 or Lotze,32 who, while they formulated the principle of the causal
theory of time quite precisely, did not enrich it with any new ideas.
Lechalas was the first to give the causal theory of time an explicit and
systematic exposition. The very title of the chapter devoted to this theory in his
Etude sur l'espace et Ie temps33 (Study of Space and Time) is in itself a
considerable fmd: 'Identite de la relation temporelle et de la relation de
causalite occasionnelle (,The Identity of the Temporal Relation with the
Relation of Occasional Causality') - which, if we leave out the last French word,
already recalls much more language of recent causal theories than that of those
which he claimed as precursors. He seems to have realized, perhaps before
anyone else, that Leibniz's relational theory does not solve all the epistemological
problems about time, that it is important to recognize that time is a set of
relations, and not a 'thing', a receptacle for events, but that the essential point
would be the ability to analyze the temporal relation, to reduce it to simpler or
more rational elements. He thought he had found in Kant the true solution to the
problem: "Kant developed the essential points of the causal theory of time, 34
and our task is to complete the considerations put forth in his second and third
analogies." 35 For Leibniz, the causal theory of time was perhaps nothing but one
more analytic definition; for Kant, it was the secondary result of the proof of the
principles of causality and universal interaction, which were his main interests,
but very difficult to fit into the system of Kantian doctrine. Lechalas recognized
the importance of the problem. He attributes its solution to Kant, but perhaps it
was necessary to have known the solution in advance in order to fmd it in the
analogies of the Critique. Furthermore) even the spirit of Lechalas' exposition
differs com-pletely from Kant's, as we shall see in what follows. Lechalas thinks
as a positive scientist rather than as a philosopher, and his epistemology is above
all a philosophy of the sciences. He is interested in the time of physics and
psychology rather than in the time ofthe immediately given and of common
sense, which is important and irreducible to the former.

70
LECHALAS' ADAPTATION OF THE CAUSAL THEORY 71

2. THE PRINCIPLE OF DETERMINISM, CONSIDERED AS DEFINING THE


TEMPORAL ORDER

Here is how he formulates the principle of the causal theory of succession:


"If we consider the world of bodies, the principle of mechanical determinism
asserts that the state of a system of material points at a given instant is detet-
mined by its earlier states and determines its later states: for us, this principle
can be reduced to the assertion that the states of this system determine one
another, and that the determining states are said, by definition, to be earlier
than the determined states, each state being, moreover, both determined and
determining, depending on how one considers its relation to one or another
of the various states." (Etude, p. 175.)
The chief merit of this statement is that it poses the question in clear
terms and restricts it to a ftxed domain. It nevertheless raises many problems,
some of which were anticipated and analyzed by Lechalas. A complete and
exact enumeration of all the concepts involved is of paramount importance in
the study of this causal theory oftime, where the main pitfall to avoid is that
of circularity (causal relations used in the defmition of time can, in turn,
presuppose the notion of time). Thus, in the defmition proposed by Lechalas,
according to which that one of two connected states of a material
system which determines the other is the earlier, the notion of the 'state
of a material system' already seems to contain the notion of time, for several
reasons. First of all, the state of a system of material points is a group of
simultaneous states of all the points of the system; the concept would then
presuppose simultaneity, which is a temporal relation. Lechalas does maintain
that "the notion of a static state of a system of material points is, in reality, an
extra-temporal notion. Time will appear when we consider several different states
of the same system of points but none of these states, considered in isolation,
involves the idea of time." 36 But is that really the case? Obviously Lechalas sees
the instantaneous state of the system as completely determined by the mutual
distances between its points, i.e., by its geometrical aspect; thus, it would involve
only space, and not time. But if the instantaneous state of a system were
completely determined by its geometrical conftguration, we would have to
conclude that its evolution depends only on the trajectories of its points. But this
is false, since there are systems which go through the same trajectory in the
course of time, but whose evolution nevertheless differs from one case to
another, if only with respect to the speed of travel over the course. Consequently,
the notion of the state of a material system, even the notion of the state of a
material point (since a point by itself can constitute a
72 THE CAUSAL THEORY OF TIME

system), implies something more than the indication of its spatial position, and
which must also be taken into account, in its analysis of the instantaneous
tendency toward change, Gust at the moving point), otherwise known as the
Leibnizian 'conatus'.
What must then be understood by 'momentary state of a material point'?
Certainly not the aggregate of all of its properties at a given moment, for this
would imply a knowledge of the relations between this point and everything else;
nor even the aggregate of those of its properties which are physically
ascertainable at this instant, for this too would imply a knowledge of the point's
relations with the present universe. Practically speaking, by 'the momentary
physical state of a material point' we mean the statement of its position and
velocity at the -given instant. Why don't we also include the acceleration or
higher derivatives with respect to time, which are also clearly instantaneous
properties of the point? The answer is obvious: in order to be able to state the
principle of causality as simply as possible. Since the Newtonian laws of motion
are second degree differential equations, whose integrals are determined by two
initial conditions, to be able to assert that the course of an isolated material
system is completely determined by its initial state, we will have to define the
physical state by two properties. The con-ventional side of this definition is
obvious. If the differential equations of motion were of a higher degree than
second, we would define the state of the (material) system differently, in order to
keep the principle of causality in its usual form, which is in agreement with the
vague but nevertheless tenacious intuition of common sense. Must we conclude
from this that physical causality is not a relation logically definable in terms of a
space-time geometry, but rather that it is always relative to our knowledge of the
laws of nature? Or are we to conclude that the principle of causality is not
rationally deducible from the spatio-temporal description of phenomena but is a
result of the special form of the laws of classical physics and of a conveniently
chosen definition of 'physical state'? We think not; what is conventional in the
definition of the instantaneous state of a physical point is that, for reasons of
economy, we take into account only two instantaneous properties. The choice of
these two particular properties, spatial position and velocity, is also a
conventional one. These are chosen because they seem to be more easily
manipulable than other theoretically possible ones (e.g., the instantaneous values
of derivatives of a higher degree). But what is not conventional is the fact that the
knowledge of the position and of the instantaneous velocity (or of two other
suitably chosen properties) is sufficient to determine the evolution of the system
in a unique manner. We conclude from this that the vicious circle in the causal
definition
LECHALAS' ADAPTATION OF THE CAUSAL THEORY 73

of the succession of physical states is only apparent. True, the usual


definition of physical state contains the notion of velocity, which in turn
contains time explicitly. But we can avoid the vicious circle if we notice that
this definition is connected with a special method of determining the
physical state of a material point in terms of its position and velocity relative
to a system of coordinates: this state can also be described by means of other
parameters, whose determination no longer explicitly presupposes
knowledge of tempo~al facts (e.g., the instantaneous value of the energy,
defined in turn by the thermal equivalent, or the instantaneous value of the
force applied to the point). A distinction must be made here: the state ofthe
material point could be determined by parameters theoretically defmable
without recourse to the notion of time, although time occurs in every
physical experiment. We shall return later to this important distinction
(Chapter V, Section 3), accepting, with the caution which it suggests, the
possibility of an extra-temporal defmi-tion of the state of a material point.
If we admit that the state of a material point is perfectly determined by
properties whose defmitions do not explicitly involve time, the fact still
remains that in order to choose from the sequence of states of each point that
state which is a member of a particular group constituting a state ofthe
system, a principle of choice must be employed which is none other than
simultaneity. Perhaps a schematic form of reasoning will help to point out the
need for a principle of choice. Let us consider a system S containing n
material points Pl. p z • ... ,Pn , each one of which passes through a series of
states (continuous, as in classical physics, or discontinuous, as Lechalas
supposed, in this respect anticipating quantum physics).37 Let us call E~ the
m-th state of the point Pn (the variable m shall be continuous or discontinu-ous,
as the case may be), and let us arrange the states of each point in some arbitrary
extra-temporal order. The state of the system will then be repre-sented by a group
of states (ETI, E'r 2, ... ,EWn). How will the physicist, in order to complete the
state to which ET I belongs, choose from the set E2 the element E'r2 rather than
any other, or in general, the state E'/(k from the group Ek (1 < k ~ n)? Clearly
by determining which states are simul-taneous with ETI. The notion of the state
of a system of material points does, therefore, presuppose the notion of
simultaneity, and the same is true of the principle of mechanical determinism
employed by Lechalas. Conse-quently, in order to complete his theory, it would
be necessary to defme simultaneity independently of the principle of
determinism. To do this, we need only follow in Kant's footsteps and defme
simultaneity in terms of interaction. In Newtonian physics, universal gravitation
constituted an
74 THE CAUSAL THEORY OF TIME

interaction precisely in the Kantian sense and could therefore serve to define
simultaneity. For example, to decide which position of point Pz is simul-
taneous with some position of point PI, one would only have to measure the
force with which PI attracts Pz when PI is in the stateETI [and call it F:]
among all the states E z through which the point Pz passes, the one which is
simultaneous to ETI will be that where the force exerted by [Pz on PI] will be
exactly equal and opposite [to F] . It is easy to imagine several other
procedures allowing us to deduce a definition of simultaneity from the law of
gravitation, and it is clear that the instantaneous propagation of gravitational
force must figure either implicitly or explicitly in all of these defmitions.
In thus completing Lechalas' theory to conform with Kant's we would
only be replacing the principles of causality and universal interaction in
Kant's definitions of succession and simultaneity with statements of a more
precise physical meaning - the principles of mechanical determinism and
universal gravitation. Later, we will try to analyze the significance of this
procedure, which consists in taking a basic scientific principle as a defmition,
and which, thanks to Poincare especially, has become so important in
contemporary philosophy of science. For the moment, we will limit our
attention to fmding out whether Kant's two principles, or the physical
equivalents which we just substituted for them, are equally necessary for the
definition of a temporal order of becoming. Let us note, in fact, that the law
of universal gravitation was unique in pre-Einsteinian physics, since the
aggregate of non-gravitational phenomena was governed by laws which
presupposed no instantaneous propagation. Universal interaction was not
defined for these phenomena, whose process of change nevertheless revealed
the same temporal order as that of gravitational phenomena. This order was
definable independently of the notion of interaction, and one suspects that the
principle of determinism, alone supposed valid, was adequate for its
definition. Indeed, we will show that by pushing the causal explanation
further than Lechalas did, we can deduce the temporal order of succession
and simultaneity from the principle of mechanical determinism.
Let us call a 'configuration of the system' any set of states containing one
and only one state of each point of the system 38 (a state of the system is
therefore a configuration of simultaneous elements). Any decomposition of
the set of states of the system into configurations will be called a 'causal
decomposition' if it satisfies the following conditions: (1) every element of
the set is part of a configuration, (2) two different configurations have no
element in common, (3) these configurations form a causal series, i.e., are so
ordered that each preceding configuration determines the subsequent one (we
LECHALAS' ADAPTATION OF THE CAUSAL THEORY 75

will return later to the important notion of determination). We can see that
the only effect of introducing the temporal order into a mechanical system is
to divide the set of the states of the points of the system into different classes,
which can be considered as the configurations resulting from a causal
decom-position of the set E~. Now, the properties (1)~(3), characteristic of
causal decompositions, contain neither the concept of simultaneity nor that of
succession: they can therefore be used to define both. In other words: the
events of the system must be decomposed into 'simultaneous' and 'successive'
events so that the principle of determinism is satisfied, by definition. Thus,
to decide whether, of the two states E'J: and Em,', one is prior to the. other,
n .
[and if so, which one], or whether they are simultaneous, all the states of
the puints of the system must be arranged into configurations satisfying
conditions (1)-(3). If the two events are in the same configuration, they are
said to be simultaneous; if they belong to two different configurations, the
one belonging to the determining configuration will be said to be prior to the
one belonging to the determined configuration. The configurations which
constitute a causal decomposition are the states of the system; the temporal
order of the states coincides with the causal order of the configurations. In
Newtonian physics, the only physics envisaged by LechaJas, for any given
system there was only one decomposition satisfying conditions (1)-(3).
Special Relativity has shown that for a given material system, there is an
infinite number of such decompositions, each one relative to a choice of
coordinate system. The properties of the temporal order can be expressed in
terms of those of causal decompositions; for example, the absolute character
of simultaneity corresponds to the uniqueness of the decomposition. The
advantage of this method of causal decompositions over the combined
method of Kant and Lechalas discussed above is that it does not make use of
the principle of universal interaction and is therefore capable of
generalization in case science were to reject the hypothesis of instantaneous
propagation while still keeping the principle of determinism. 39
But here is a new difficulty arising from the use of this principle: for a
state of a system of material points to determine its subsequent states, these
must depend only on the forces within the system, or more generally, the
variation of the field of forces in which the system evolves must be fixed in
advance; it is clear that in different fields, the same initial state ofthe system
could lead to different subsequent states. The principle of mechanical de-
terminism invoked by Lechalas can therefore be stated more rigorously as
follows: "The state of a system of material points at a given instant t is
determined by the state of the same system at a prior instant to and by the
76 THE CAUSAL THEORY OF TIME

variation in the field of forces to which the system is subject during the
interval separating these instants." In the particular case of an isolated
system, all of whose forces emanate from its own points, the principle of
mechanical determinism can be formulated more simply: The final state of a
system of material points is determined by its initial state, provided the
system remains isolated during the time interval separating these two states.
Thus, in both cases, the principle of determinism involves not only the
determining and determined states, but <hSO presupposes a certain knowledge
of what takes place after the determining state and before the determined
state: the notion of before and after, explicitly contained in the principle of
determinism, could therefore not be denied in terms of it. Lechalas'
definition seems plausible only because of an incomplete, and therefore
incorrect, formulation of the principle of mechanical determinism.
We arrive at the same conclusion if we consider the more general
principle of physical determinism instead of the principle of mechanical
determinism. It is important to note here that the very restricted formulation of
deter-minism, cited by Lechalas (whereby the only process whose causal nature
is asserted is the motion of a system of material points under the influence of
internal forces) can be extended to the whole of classical physics, and also, in
a certain sense, to quantum mechanics. Here is P. Jordan's formulation of
this generalized principle in his noteworthy discussion of 'causality and
statistics in modern physics'.

Classical field physics means that physical reality can be described ... : for every point in
a four-dimensional time-space framework certain measurable quantities - field strength,
gravitational potential, etc. - are stated numerically. Hence a causality exists in the
following sense: let us imagine a finite portion of space, say, in the form of a box ... At a
fixed time - let's say 11 o'clock - let the physical state inside the box be measured
completely. In addition the physical state of the entire surface of the box should be
controlled between 11 and 12 o'clock. In such circumstances the physical processes
inside the box are unequivocally determined. If the initial state of the box and the
temporal course of processes on its surface are reproduced at any time in any place, all the box's
4o
inner processes are also reproduced.
Here is the sense in which this principle of determinism of classical
physics still survives in quantum mechanics:

Quantum mechanics describes the world by means of an abstract coordinate space of


enormously many dimensions: the number of dimensions is proportional to the number of
all the available material particles in the world. In this abstract space extended entities
move constantly, which do not directly describe each single event in the world of atomic
phenomena but only determine the probability of quantum-like processes. Causality -
LECHALAS' ADAPTATION OF THE CAUSAL THEORY 77

conceived not as a metaphysical antithesis to a metaphysical concept of chance, but rather


understood as the previously formulated physical statement - is valid formally in quite the
same way for both theories.41

Jordan's formulation clearly contains Lechalas' theory as a special case: the


transition from one to the other presupposes that the 'initial conditions' of the
field within the box are reducible to the distribution of the masses and velocities,
and that the 'boundary conditions' regarding what takes place on the surface of
the box require only that no physical action be exerted through the walls of the
box (the postulate of an isolated system). It is these boundary conditions which
implicitly contain the notion of before and after, since they must be realized in
the time interval separating the initial state from the final state, i.e., after the
former and before the latter. We return then to the conclusion that one cannot use
the principle of physical determinism in a causal defmition of succession without
involving oneself in a vicious circle. Notice, however, that the knowledge of the
boundary conditions (and, in par-ticular, that of the isolation of the system)
becomes superfluous if we apply the principle of determinism to the entire
universe. In this case, Lechalas' definition would be: of two states of the universe,
the one which determines the other is said to precede it in time. Here, our
objection, that the notion of a state of the universe (which is an instantaneous
cross-section of the universal becoming, formed of simultaneous events)
presupposes the notion of simul-taneity, would be valid, and its avoidance would
require the application of the method of causal decompositions. To formulate
these results more simply, let us suppose that an extraordinary being, who
simultaneously perceives all of the events which constitute the history of the
universe, those now belong-ing to the past, as well as those to come,42 noticed
that men arrange these events in an order which they call temporal, and decided
to fmd the principle governing this order. Obviously, he will never succeed in re-
inventing the human method, which uses the temporal order of perceptions to
establish the temporal order of events, since he simultaneously perceives the
cosmic evolution (although not his own). But he will be able to invent another,
which will lead to an ordering of events identical to that made by man. (That two
distinct methods should lead to the same ordering of a set is not surprising: for
example, one arrives at the same ordering of whole numbers either by comparing
them directly with one another or by taking them as exponents and comparing the
resulting powers of some number greater than one.) For this purpose he will be
able to use the causal relation, which he is supposed to know directly. He might
say the following: "I will try to separate
78 THE CAUSAL THEORY OF TIME

all events into classes with the following properties: (1) each event is a mem-ber
of one and only one class, (2) given any two classes, one is always the cause of
the other (and is said to be earlier than the other). I will call the decomposition of
events into classes satisfying these two conditions a causal decomposition, and
the classes themselves the instantaneous states of the universe. 43 There are
an infinite number of causal decompositions, and it is very surprising that until
1905 everyone believed that only one was possible. Two events belonging to the
same state of the universe are said to be simul-taneous relative to the temporal
decomposition to which this state belongs; event X is said to be earlier than event
Y relative to a causal decomposition if this decomposition contains two states of
the universe such that the one containing X is the cause of the one containing Y."

This is the way that the superhuman representative of the causal theory of
time would speak, and we would not dare accuse him of circularity in his
definitions. But doesn't the fact that it is being professed by a superhuman being
clearly point to the metaphysical character of this theory? In fact, the extension
of the principle of determinism to the whole universe seems to be an essentially
unverifiable hypothesis, hence metaphysical par excellence, and the same is
true of any theory which rests on this hypothesis. Besides, it is certainly in a
metaphysical sense that Lechalas interpreted the causal theory of time. To this the
superhuman representative of the causal theory of time would answer:
"Undoubtedly, the extension of determinism to the whole universe is not directly
verifiable by human means and therefore belongs to metaphysics, if you want to
call metaphysical any hypothesis which, humanly speaking, does not admit of
direct verification. But isn't the same true of the hypothesis of the temporal
order? This one amounts to saying that all events of the universe can be arranged
in the order of succession and simultaneity so as never to put the same event in
two cross-sections of simultaneity, nor effect before cause. It too is therefore
directly unverifiable, just as much as the hypothesis of universal determinism,
and the causal theory of time has the right to replace one by the other."

3. PHYSICAL REVERSIBILITY

But this theory,in spite of becoming metaphysical,is not beyond questioning. In


Lechalas' definition, according to which one state, which is the cause of another,
is prior to it, the term 'cause' undoubtedly requires a detailed analy-sis. Lechalas
says:
LECHALAS' ADAPTATION OF THE CAUSAL THEORY 79

For Hume, the cause, or better still, the set of causes, is the totality of conditions pre-ceding the
production of a phenomenon; for us, on the other hand, in a group of facts, those which are the
condition of others are said to precede them, and the second follow the first, without these
expressions signifying anything other than this relation of 'occa-sional causality,' to use
Malebranche's terminology.44

Thus, the later event could not occur without the earlier event, which is the
condition for it. But isn't the converse also the case? Could the earlier state
exist without being followed by the later state which is thereby also its condi-
tion? This seems obvious, if we remember that the set of causes is not only a
necessary but also a sufficient condition for the effect, the latter then being a
necessary condition for the cause: to say that A is a sufficient condition for B
is the same as saying that B is a necessary condition for A. Essentially, the
determination of the states of the universe by one another amounts to this:
the production of state El at time t 1 is the necessary and sufficient condition
for the production of state E2 at time t2 , and vice versa (or, in the particular
case considered by Lechalas: by fixing the initial conditions for a system of
material points, you determine not only its future, but also its entire past).
The only condition which the numbers t 1 and t 2 must satisfy is that t 1 =1= t 2 •
Physical determination is therefore a symmetrical relation and cannot serve
to defme the asymmetrical relation of before and after.
One might reply that determination, symmetrical in abstracto, becomes
asymmetrical in particular physical laws. These laws not only link phenomena
symmetrically in time, but also determine the temporal order of phenomena, and
can be used to define it. To know which of two mutually determined phenomena
precedes the other, we can consult the law governing this deter-mination: it will
tell us which is the cause, and consequently, which is earlier. But is this true?
Does physics really teach asymmetrical causality? We are touching here on a
problem which was famous at the time of the publication of Lechalas' book but
which is [surely] very obscure today, concerning the reversibility of physical
processes. The elementary mechanical processes are all 'reversible,' they can take
place in both directions. Let us consider, for example, the free fall of a heavy
body endowed with an initial velocity v; under the influence of the earth's
gravitational field, this body after a certain time will reach a new position and a
fmal velocity v'. If we threw the same body from its fmal position with an initial
velocity equal but opposite to v', obviously it would pass through the same series
of points with the same velocities changed only in sign, and, in its fmal position,
would reach the velocity (-v). Thus, the two phenomena of'the fall and ascent of
the heavy body are perfectly and mutually symmetrical with respect to time.
From a
80 THE CAUSAL THEORY OF TIME

study of their intrinsic properties we would be incapable of deciding which of


the two is a fall or an ascent. Certainly there is physical determination be-tween
the initial and fInal states of each phenomenon, but it seems impossible to decide
which state precedes the other by only studying the phenomenon itself. To use
the language of Leibniz, the physical present seems pregnant with the past as
well as with the future. We can see that this objection con-cerning the
reversibility of physical phenomena, raised by Lechalas, is of capital importance
for the causal theory of time. It did not escape Bertrand Russell, who summed up
the problem thus:

Can time be derived from causality, or must we retain temporal order as fundamental and
distinguish cause and effect as the earlier and later parts in a causal relation? The question is
bound up with that as to reversibility of physical processes. If causal relations are symmetrical
so that whenever A and B are related as cause and effect it is physically possible that on another
occasion, B and A may be so related, then we must regard the time order as something
additional to the causal relation, not derivative from it. If, on the other hand, causal laws are
irreversible, then we define the time order in terms of them, and need not introduce it as a
logically separate factor. The question on rever-sibility is still sub judice, and I will not venture
an opinion ... ,,45

Certain distinctions can still be made here; in reality the question is not so simple
as we might be led to believe by the lucidity of these lines, but de-mands a
thorough examination. We must not confuse the reversibility of physical
processes with the symmetry of the causal relation. If a physical process,
decomposable into consecutive phases A, B, C, where A is the cause of Band B
the cause of C, could occur in reverse, B becoming the cause of A, we should
expect that also in the fIrst process B would be followed by A and not by C, and
that therefore the cyclic process ABA would be repeated indefinitely. If we
substitute a continuous change for the sequence ABC, and suppose it to be
reversible, we can apply the same reasoning to any phase of it, as close to the
initial state as desired, and conclude that these two states will repeat themselves
indefinitely. This shows that no causal process (i.e., such that of two consecutive
phases, one is always the cause of the other) can be reversible, if we mean by that
the possibility of passing in the opposite direction through the same series of
states.46 In the examples of mechanical reversibility, a body passes for a second
time through the same positions, but its velocity changes in sign. Since a
momentary state consists of a velocity as well as a position, it would be incorrect
to say that the body passed for a second time through the same states. It is not the
direction of the process that we change, but the direction of time itself.

A spatial analogy strongly suggests this idea: in studying the position of a


LECHALAS' ADAPTATION OF THE CAUSAL THEORY 81

curve with respect to an axis, we note the same change whether we turn the
curve or change the direction of the axis. However, even in the spatial image, if
we attribute to each point not only a distance, but also a direction of the tangent,
'reversibility' is no longer symmetrical. We then see that a point traversing a
curve symmetrical with a given curve does not go through the same series of
states in the opposite direction, although it does go through the same series of
positions. In the case of a temporal process, the derivatives with respect to time
are always implied in the instantaneous state (which is perhaps nothing but the
mathematical expression of the Leibnizian 'conatus'). Therefore, there cannot
be any process capable of taking place indifferently in either direction.

Thus, in the case of mechanical reversibility, it is not the process, but time
itself whose direction is changed. Here, approximately, is what must be
understood by that: suppose that the future becomes the past for you, and that the
past takes the place of the future. If you are a physicist you will notice no change
in the laws of mechanics. Processes will continue to be determined as before by
their initial conditions. As a partisan of the causal theory of time, you will have
the disagreeable surprise of deducing the oppo'lite tem-poral order between any
two phenomena, from that established by your colleague, who would not have
changed the sign on his temporal variable. This divergence between your results
and his will be all the more disturbing because you will have every reason to fear
that it is occurring within your own time. If your time were well ordered and
coherent, you might be able to consider the different temporal orders as so many
different ways of expressing the same physical reality. But, since the temporal
order which you introduce into phenomena is entirely deduced from causal laws,
and since these have led both you and your colleague to divergent orders, what
assurance have you that these same laws, applied by you twice in succession, or
in two intervals of time, will not lead you to establish two contradictory temporal
orders?
This incapacity of the causal laws of physics to determine the temporal
direction of becoming (which is translated mathematically by the absence of
derivatives of odd degree with respect to time in fundamental equations)
undoubtedly comes from the fact that the physicist has little interest in the history
of the physical universe, having decided in advance to consider events only as
instances of the laws which govern their development. But ask an astronomer or
geologist whether the temporal order of becoming seems indeterminate to him,
even whether he believes most tenaciously that all the laws of the world are
reversible. Undoubtedly, the geologist will answer you in the negative: "the Earth
is an individual, and its history is that of an
82 THE CAUSAL THEORY OF TIME

individual. Even if the evolution of the Earth in the opposite direction were
theoretically possible, this would not change the fact that the historical evolution
took a particular direction and not another. It is possible, although infinitely
improbable (he may continue) that there is a material system in some part of the
world which goes through the same series of states which constitutes the history
of the Earth, but in the opposite direction. These two Earths may very well be
interchangeable in many respects, but they are nevertheless distinct, that is they
cannot be interchangeable in every respect without contradiction.47 If they are
interchangeable for the physicist, it is because he is concerned only with certain
aspects of becoming (the proof of this being that, using his own methods, he
never succeeds in individualizing a single phenomenon.) As in the past, I shall
continue to call the era of the Earth preceding the mesozoic era, the archaic era,
even if on another Earth the order of their counterparts is reversed. In spite of the
fact that time follows reversible physical laws, it nevertheless seems to me
irreversible in itself. The appearance of the earth's history would be completely
changed if the direction of time were reversed. The same is true of space, whose
physical isotropy, at least on our scale, has never been contested. The intrinsic
pro-perties of a body do not change if the body undergoes a rotation in space.
Nevertheless, it is clear that in general the events which take place on one
straight line differ from those taking place on another. Thus, there is isotropy for
the physicist, but not for the geographer or astronomer. Similarly, for me, a
geologist, or for any historian, since the distribution of events in time is certainly
not reversible, time is not isotropic, even if it were so in the eyes of the
physicist."

However, because of its irregularity, the irreversible distribution of events in


time, as taught us by history in the widest sense of the word, cannot serve to
define the temporal order. But are there no regularly irreversible processes which
could serve as infallible indicators of succession? A gourmet would undoubtedly
answer in the affirmative: "Vin sur biere fait l'affaire, biere sur vin ne vaut rien"
[Wine after beer is a good thing, beer after wine a poor one.] There are
undoubtedly series of phenomena which occur in one direction rather than
another. The second law of thermodynamics, according to which the entropy of a
closed system always increases (or, rather, never decreases), governs an
extremely large class of such processes and dominates the whole of physical
evolution. Still, these sequences are not causal sequences at all, and a theory of
time which derived the temporal order from the course of entropy would hardly
deserve to be called a causal theory.48 Let us note, in fact, that an irreversible
process of thermodynamics, such as the transfer of heat from a
LECHALAS' ADAPTATION OF THE CAUSAL THEORY 83

warmer body to a colder one is not a causal process: the earlier state, i.e., the
initial distribution of heat, does not uniquely determine the fmal distribution. To
be in a position to assert the existence of a causal relation, it would be necessary
to know the exact distribution of positions and velocities, precisely what is
dispensed with in the second principle of thermodynamics. An irreversible
phenomenon is nevertheless an indicator of succession, and can serve to define it.
We say 'indicator of succession' and not 'clock,' since the metrical properties of
time do not enter into Clausius' principle. The causal definition of simultaneity
does not depend on the reversibility of phenomena. In Lewin's and Reichenbach's
'method of messages,' which we will explain later, messengers, i.e., causal chains,
are sent out from a material point, returning to it after having reached another
point during their travels. It is easily seen that by using fast messengers traveling
small distances, one can define the simultaneity of events taking place at two
different points with great precision. Now, this definition obviously does not
depend on the direction of time, since under either of the two possible hypotheses
on the direction of time, the same events turn out to be 'simultaneous.' By com-
bining the causal definition of simultaneity with the statistical definition of
succession, we can obtain a theory of the temporal order which is satisfactory in
certain respects. This is perhaps the sense of Lechalas' brief comments on the
compatibility of the causal theory of time with the reversibility of physical
phenomena. Basically, he renounces a homogeneous causal explana-tion of the
temporal order in favor of a semi-causal, semi-statistical theory. Is this
compromise, or even Russell's total renunciation, really necessary? By using the
'method of the causal decomposition of becoming,' could we not deduce the
temporal order of events from their causal interconnection, even supposing the
latter to be reversible? We shall try to show that this is indeed the case with
Reichenbach's work, where we shall fmd a systematic exposition of the statistical
theory of succession, and we shall then see the serious dif-ficulties which arise
from this theory.

4. PSYCHO-PHYSIOLOGICAL REVERSIBILITY

Lechalas tried to refute the objection of reversibility in another way. He says that
even if physical causality were symmetrical, and allowed no way to distinguish
the past from the future, "essentially, the interconnection of physical phenomena
can take place in only one direction, and this direction determines that of spatial
phenomena, because of the link between the two orders of phenomena: a pin-
prick precedes pain, since it causes pain, and,
84 THE CAUSAL THEORY OF TIME

similarly, the will to raise my arm precedes this movement."49 Lechalas'


reply raises two difficulties: first of all, it is not clear how the supposedly
irreversible chain of psychic facts can be extended to the entire material
universe; only a physical event which is the cause or effect of a psychic fact
would thereby be situated in time by its relation to this fact. The temporal
relations among physical events themselves would not be affected. Secondly,
the very nature of the temporal interconnection of psychic facts, insofar as it
is a new given with respect to the irreversibility of physical phenomena,
requires detailed analysis. Is the temporal distribution of states in a psychic
flux really irreversible, as it appears on fust sight? It undoubtedly is in many
cases; for example, it seems impossible that the memory of an object is prior
to its perception, or that hope follows the realization of a desire. But there
also appear to be psychic phenomena which can sometimes take place in one
direction, sometimes in another - such as the different perceptions during the
displacement of a picture in two opposing directions. However, this
reversibility could be only apparent, if, as some seem to believe, it is true
that every psychic state in some sense reflects its immediate past,50 or even
its entire past.51 Whatever the case, the sequences of psychic states which
we mention fust do seem irreversible. It is important to know whether this is
a new fact, characteristic of the psychic life, inseparably linked to all
spiritual becoming, or rather a derivative phenomenon, explainable in terms
of an irreversible physical process. Only in the fust case would we have a
new definition of succession.
We must distinguish here between the reversibility of a group of phenomena
within a given psychic flux, and the reversibility of the whole flux. This last
might be impossible if the psychic flux formed an organic whole whose parts
were interconnected. But the whole psychic flux might still be reversible, as
Plato already conjectured. Notice, however, that the possibility of this apparently
absurd hypothesis results from two principles which have been shared by
scholars for a long time, that of universal mechanism in physical processes, and
that of psycho-physical parallelism in the relation between the physical and the
mental. Suppose that all physical becoming is reducible to motions governed by
reversible laws, and further, that the order and nature of the states whose
succession constitutes some arbitrary psychic process are completely determined
by the corresponding physiological process. In par-ticular, physiological
processes taking place in reverse would correspond to reversed psychological
processes. At a given instant a man, having reached middle-age, would be
represented physically by a configuration of molecules each with a determinate
velocity at that instant. According to these two
LECHALAS' ADAPTATION OF THE CAUSAL THEORY 85

hypotheses of universal mechanism and psycho-physiological parallelism, a 'new


Medea' 52 wishing to rejuvenate this man would only have to change to signs of
all these velocities simultaneously. She would see him retracing his steps in time,
undergoing the series of states in reverse order which consti-tuted his life until
the moment of rejuvenation - provided, however, that the external conditions
were favorable. This is not an insunnountable difficulty, since it would be
sufficient to arrange things so that the immediate environ-ment of the new Aeson
would be practically constant during a certain interval of time before and after
the rejuvenation. Admittedly, his feeding would present an enormous difficulty
for Medea if she wanted to care for her father-in-law herself. But he could
probably get along by himself, provided we had thought of accumulating a large
enough stock of food of stable composition in his immediate environment (the
stability of the composition and the size of the stock are required .to assure the
approximate invariance of his surround-ings). The rejuvenation by simple
temporal reversal would be of little practical interest. No doubt Medea would
think of markedly accelerating the reverse process and, after some time, of
making a second change in the velocity of the molecules constituting the body of
her father-in-law. She would then perceive that he had started to behave
reasonably like other mortals; in particular, had begun to express himself
intelligibly, which had not been the case in the interval between the two changes
of velocity. In order to understand Aeson's words during that interval, the new
Medea had been forced to record them and play the recording backwards. It is
true that this tiresome procedure made Aeson's words clearer and very
reassuring: t seconds after the first change in velocity, he would only repeat what
he had said t seconds before this change. It was obvious that his states of mind
after this change were also a repetition, in reverse order, of his states of mind just
before the change. This would not surprise Medea: she had thought about this for
a long time in connection with certain film makers whose comic imagination
could fmd nothing better to do than to run the latest banal fIlm backwards. But in
Aeson's case there was the realization of cinematographic buffoonery in a
psycho-physiological sense as well as the enormous practical interest of his
metamorphosis. Aeson's impressions during the critical interval would have been
similar to Medea's. He too would see in what went on around him a movie played
in reverse, although correct in his particular case. It is even possible that Aeson
would refuse a new change of velocity, preferring to live in this new world in reverse.
In an environment suitably managed by Medea, he would no doubt succeed in getting
along after some trial and error. Perhaps he would try to explain his difficulties by
recalling that he lived in a
86 THE CAUSAL THEORY OF TIME

cosmic period of very low statistical probability, dominated by irreversibility


as a result, i.e., by the continuous increase of this probability. If, he would
tell himself, the vital processes such as they appear in the actual organic
world are in principle irreversible, it is perhaps because there is here an
adaptation of life to our cosmic period. In a period where entropy, i.e.,
statistical pro-bability, reaches a maximal stable value, his new temporal
orientation, strange at present, would no longer be so.
We believe, then, that there is nothing absurd about the hypothesis of a
psychic flux evolving in the opposite direction from ours. Although they run
counter to our ingrained habits, Aeson's impressions during his period of
rejuvenation as well as those of a spectator watching a movie being run back-
wards and so immersed in the spectacle that the rest of his psychic life is
almost effaced, involve no contradictions. Besides, there is no need to appeal
to the hypothesis of psycho-physiological parallelism to link the approximate
and vague irreversibility of psychic phenomena with the statistically mea-
surable and somehow rational irreversibility of physical becoming. Even if,
as Bergson would have it, psychic facts were only 'attached' to cerebral
processes, (which seems to be the minimal hypothesis about the psycho-
physiological relation), and, by their novelty or unexpectedness, went com-
pletely beyond what occurs in the nervous system, the assumption of the
irreversibility of the physiological series would still force us to infer the
irreversibility of the psychic one, since any given psychological fact cannot
be 'attached' to just any cerebral state. Furthermore there seems to be no
reason to believe that a psychological fact attached to a given cerebral state
could also be attached to the 'inverse' state (the state resulting from a
simultaneous change in sign of all the velocities). Thus, Lechalas' defmition
of succession in terms of psychological irreversibility does not seem to
extend our knowledge of objective time, nor does it seem to be suitable for
giving a new explanation of it.
However, it might be objected that psycho-physiological; or, more gener-
ally, biological irreversibility is a fact which has never been contradicted by
experience, while the irreversibility of inorganic processes is only a statistical
law admitting of exceptions. Thermodynamic fluctuations, the object of
much study since the well-known work of Einstein and Smoluchowski on
Brownian movement, seem to prove that in certain cases heat can pass spon-
taneously from a cooler to a warmer body, and that in other cases, a certain
quantity of heat can be entirely transformed into kinetic energy. Thus, the
phenomena of heat transfer and dissipation of energy no longer seem to be
absolutely irreversible. But no one has seen a living creature who, instead of
LECHALAS' ADAPTATION OT' THE CAUSAL THEORY 87

aging with the passage of time, grew younger and younger. Is it not the case,
then, that these irreversible biological phenomena reveal an absolute
anisotropy of time?
We shall not debate the factual question of the existence of absolutely
irreversible biological laws, although one might ask, for example, whether
the recent attempts at artificial rejuvenation do not give reason to doubt the
absolute irreversibility of the law of senescence. On the contrary, let us admit
the existence of absolutely irreversible biological laws. Must we conclude
from this that time is anisotropic? We doubt it. Let us note, in fact, that
according to evolutionary theory, all ofterrestriallife must have originated at a
particular instant and at a particular point on the surface of the globe. For
some reason or other, a group of molecules, each with its own determinate
velocity, happened then and there to form the first piece of living matter. This
fact by itself is sufficient to explain biological irreversibility. There is no
need to insist on the uniqueness of this first being, nor to affirm that an 'elan
vital,' which enveloped it, must have persisted in all its descendants, imprinting
upon them a common direction in time. It is enough to note that to attribute an
isotropic structure to physical time is not to deny that a material system, placed in
a determinate setting and given a set of initial conditions, must go through a
certain series of changes in one direction, arbitrarily chosen in time, and not in
the opposite direction. What this isotropy does imply is that two identical
material systems, placed in initial conditions symmetrical with respect to time,
must evolve in opposite directions. There-fore, for the existence of reversible
biological phenomena to be necessary, it would be necessary for life to have had
its origin in two identical beings, placed in exactly symmetrical conditions to
each other with respect to time: it would be necessary, in particular, that each
molecule of one of these beings should have had the same initial velocity,
opposite in sign, as the corresponding molecule in the other. Only then would we
be assured of finding reversibility in life. It is clear that the probability of this
hypothesis is practically nil, perhaps even theoretically, since it would seem that
two beings with different temporal orientations would not both be viable under
the same terrestrial conditions.

In conclusion, then, let us say that biological irreversibility has only a


historical character, just as the anisotropy of space on our scale is only
geographical. (Yet, biological irreversibility must be qualified in one essential
way: the simple, decisive factor of biological evolution viz., the set of genetic
variations, is known to be time-reversible, [H. M., 1977].) One who derived an
absolute anisotropy of time from biological irreversibility would be like
88 THE CAUSAL THEORY OF TIME

the geographer who affIrmed that there is an absolute spatial difference between
high and low, the objects inside and outside the Earth's crust being, by defmition,
high and low, respectively. It is clear that the possibility of this geographical
defmition does not affect the physical isotropy of space in any way: similarly,
biological irreversibility in no way entails the physical anistropy of time. Let us
assume, therefore, in what follows, although we are not yet in possession of all
our arguments, that if physical laws are reversible, it is because physical time is
isotropic. It behaves exactly like a geometrical straight line. In order to defme the
concepts 'to the left' and 'to the right' for
a straight line, it is necessary to choose two points arbitrarily and agree which
one is to be considered 'to the left' of the other. The intrinsic order of the points
on the line is not affected by this convention: if the line is composed of physical
points, the distinction of direction on it will be only geographical. Similarly, to
distinguished the physical future from the past, it is necessary to choose two
nonsimultaneous events and decide which of them will be considered earlier than
the other: the intrinsic temporal order of events is not affected by this convention,
which is historical, and not physical in nature. This idea of an isotropic time
undoubtedly seems strange at first sight, just as the isotropy of a straight line
would be inconceivable for a conscious point which was attached to the line, and
would always move on it in the same direction. But when it perceived that other
objects moved on the line, but in the opposite direction, it would cease to
consider the line as a 'one way street,' in other words, it would recognize the
isotropy. We will arrive at the same conclusion later in this essay by pushing the
causal analysis of time a little further: physical time is also not a 'one way street,'
it is isotropic. This is why the efforts of Lechalas and his successors, who tried to
distinguish the past from the future by means other than those furnished by
history (in the widest sense of the word), were bound to fail. In our opinion, the
causal theory of time must set forth the causal explanation of the isotropic
temporal order, leaving to history the task: of choosing between the two temporal
directions of becoming.

5. THE EPISTEMOLOGICAL INTERPRETATION OF THE


CAUSAL THEORY

On one essential point, namely in his epistemological interpretation, Lechalas is


again unfaithful to the Kantian theory, whose main lines he claimed to follow.
For Kant, the causal theory of time is the theory of the objectification of time by
causality; but in Kant, 'objective' does not oppose 'phenomenal';
LECHALAS' ADAPTATION OF THE CAUSAL THEORY 89

it is only an immanent cognitive relation characterizing certain groups of


representations. For Lechalas, on the other hand, causal time is the true, real
time, existing outside of representations, as opposed to the felt and imagined
time, which is only its subjective spectre. Like Locke, he makes a feature of
our representation correspond to the causal structure of reality, emphasizing
that correspondence does not mean identity. He differs from Locke in that
Locke believed that besides the secondary qualities, there are the 'primary'
qualities (of which time is one), for which correspondence becomes identity;
for Lechalas, all of the qualities of representation become 'secondary,'
agreeing on this point with Berkeley, but differing from him by asserting (as
a realist) that there is a trait-for-trait correspondence between the completely
subjec-tive sensible world and the real trans-subjective universe, the universe
of science.
We can truthfully say' with Kant that time is a form of our sensibility, but a form which
conceals a rational distinction. Admittedly, this theory contradicts and renders illusory the
almost ineradicable idea which we have of time, but is this not true of any serious theory
concerning our sensibility? Consider the revolt of common sense when someone denies
that there is anything in bodies resembling our sensations of color, or that our pains are
localized in the different parts of our organism. Even greater, if possible, is our instinctual
repugnance at admitting that we really know only our states of mind, and consequently,
that the external world is merely inferred, legitimately or not. Well, in the present case, it
seems that we must still more brutally offend what we might call the intimate constitution
of our psychic life, for there is not one of our states of conscious-ness which does not fit
into the order of time and take on the temporal form, which, according to us, is illusory. If
this is the case, we should not be surprised by the revulsions which we feel in ourselves, for we
have long been prepared by similar revulsions, long since recognized to be without rational
foundation. 53

Thus, the causal theory of time, clearly phenomenalist in Kant, becomes


realist and even metaphysical in his interpreter (Lechalas does not hesitate to
use it in discussing the relation between God and the universe). And yet, after
all, this is not surprising. Most scientific and philosophical theories admit of
two interpretations, idealist and realist, with all the nuances which these
terms involve. To use the language of mathematicians, it is as if these theories
were invariant with respect to epistemological transformations, which, while
involving changes in terminology, still leave the fundamentals of the theory
unchanged. We shall have occasion to return to the metaphysical
interpretation of the causal theory of time in connection with Carnap.
Note for the moment that Lechalas is somewhat hasty in declaring that the
tempor;i1 form taken on by our states of consciousness is illusory. His
general thought is clear: just as electromagnetic disturbances are supposed to
90 THE CAUSAL THEORY OF TIME

correspond to the sensation of color, and molecular agitation to the sensation of


heat, the causal relation is supposed to correspond to intuitive succession. Yet
science, far from denying the psychological reality of visual or thermal
sensations, affirms it, in order to infer from these the structure of the corre-
sponding physical phenomena; the very validity of this inference presupposes the
reality ofthe psychological data. It seems, then, that if we infer the causal
structure of becoming from the intuitively perceived temporal structure of our
states of consciousness, we cannot declare this latter illusory without casting
doubt on the legitimacy of our inference; moreover, this is how we must
understand, mutatis mutandis (i.e., after a realist transformation) Kant's
distinction between subjective and objective time. It seems that Lechalas
attributes a causal nature not only to the objective time of physics but also to the
subjective time of psychology. But this extension is not justified in his
exposition. We have seen in fact that he derived his causal theory of time from
the principle of mechanical determinism: but this principle clearly does not apply
to psychological life. We do not deny that a thorough study of subjective time
might lead to attributing a certain ideality to it. However, such a study is not
found in Lechalas' work, which is devoted almost entirely to the analysis of the
time (and space) of physics.
From this work, with its important position in the evolution of the causal
theory of time as we have already mentioned, let us retain its concern with a
more detailed agreement with science than was achieved by the Kantian theory,
its impotence before the problem of reversibility, and the gaps in its
epistemological interpretation, arising chiefly from the almost exclusive study of
the physical form of the temporal order: whether virtues or defects, we shall find
all of these in Lechalas' successors.
IV

THE RELATIVISTIC PHASE OF THE CAUSAL THEORY OF TIME:


THE AXIOMATIC SYSTEMS OF ROBB AND CARNAP

1. ROBB'S SYSTEM

The year 1905 was a decisive one for the causal theory oftime. Until then it
was only a metaphysical speculation or an epistemological interpretation; in
1905 it directly rejoined science in its immediate consequences. The Special
Theory of Relativity demonstrated that two events which appear to be
simultaneous in one system of reference are, in general, not simultaneous in
another. If one event precedes another in one system, it could very well
follow it in another. There are many events which are simultaneous in every
admissible system, but this invariant simultaneity presupposes the spatio-
temporal coincidence of two events and cannot exist for events not meeting
this condition. Nevertheless, there is an important difference between simul-
taneity and succession (of widely separated events): simultaneity is always
relative, but thete is, in a certain sense, an absolute succession; an event
which precedes another in a given system of reference, and is its cause,
precedes it in every admissible system. And vice versa: if an event precedes
another in every admissible system, the causal relation between them
becomes possible. Hence the invariant succession relation coincides with the
causal relation. This consequence of relativity seems to have determined the
entire recent phase of the causal theory of time.
It was A. A. Robb who inaugurated this relativistic phase. Although he refers
more to Larmor and Lorentz than to Einstein and Minkowski, it is the ideas of
these last two, especially the idea of space-time as a four-dimensional physical
continuum, which dominate his great treatise. 54 Robb's work springs from
physical geometry, rather than from the theory of knowledge; he has been
nicknamed, not without reason, 'the Euclid of Relativity.' His 21 axioms and 206
theorems form the most complete and most rigorous exposition of what the
Special Theory of Relativity has to say about space and time, as expressed in
terms of a single basic concept, that of 'conic order.' Guided by a geometrical
analogy, which we will discuss later, this is how Robb designates the type of
causal order existing among instantaneous events. If £1 and £2 are any two
instantaneous events whose causal relation we are considering, three mutually
exclusive alternatives appear possible:
91
92 THE CAUSAL THEORY OF TIME

(1) Someone acting on £ 1 can produce an effect on £2 .


(2) Someone acting on £2 can produce an effect on £1 .
(3) It is impossible for anyone acting on £1 to produce an effect on £2 ,
and vice versa.

In the first case, we will say that £1 takes place before £2 ; in the second case,
£ 1 will be regarded as later than £2; in the third case £1 comes neither before
nor after £2. The relation of before and after thus dermed is characterized by
the following axioms: 55

(1) If event A is after event B, B is not after A, but before it.


(2) For any event A, there is an event which is before A and an event
which is after A .
(3) If an event A is after B, there is an event which is both after A
and before B.
(4) If A is after B, and B is after C, then A is also after C. Neverthe-
less, A may be neither before nor after B without being identical
toB.
This type of order can be illustrated by the following simple spatial model:
let us consider in Euclidean space a family of cones with parallel axes and equal
vertical angles, each point in space being the vertex of two cones; let us suppose
further that we have arbitrarily marked the two directions of the axes (the upper
and lower cones of each point). We will agree to say that any point A inside or on
the surface of the upper (lower) cone of a point B is before (after) point B. It is
easily seen that the relation of before and after thus dermed satisfies the four
axioms above. This is why Robb calls the general type of order characterized by
these four axioms 'conic.' As we have seen, he identifies the causal order of
events, which is 'conic,' with the temporal order of before and after, which
amounts to speaking of succession only where there is an invariant succession. It
seems to us, however, that no one would hesitate to say that an event which
seems prior to another to every observer attached to a given system of
coordinates, is prior to it in that system. This is the usual sense of priority and
posteriority, while that of invariant succession, adopted by Robb and other
partisans of the causal theory of time, concerns a different relation, which is
narrower and directly dermed by the causal relation. This more restricted use of a
common term only obscures the Einsteinian discovery of the relativity of
succession. - Besides, Robb does the same with simultaneity, refusing to call two
events simultaneous unless they coincide in space and time. In this way he
avoids
THE AXIOMATIC SYSTEMS OF ROBB AND CARNAP 93

speaking of relative simultaneity, which he believes to be incompatible with


the principle of contradiction. This principle would imply that it is
impossible for an object to be A and not -A, at the same time, i.e.,
simultaneously, which permits the conclusion that the relativity of
simultaneity entails the relativity of logical contradiction, which is absurd.
But this argument rests on an incorrect formulation of the principle of
contradiction, into which the concept of time has been surreptitiously
introduced. It is obvious that neither the general form of the principle of
contradiction as applied to propositions (in Russell's notation: I- 'V (p 0 0

'Vp)) nor its special form corresponding to the Aristotelian formulation: 1- 0

(x, if» o'V(if>x 'Vif», contains the notion of simultaneity; it seems to


0

because ordinary language does not contain a special symbol to designate


'logical product'. But it is sufficient to replace the logical product by some
equivalent function, composed, for example, of alternation and implication
signs, to show the non-temporal nature of the formulation. The principle of
contradiction, therefore, in no way favors absolute, i.e., invariant,
simultaneity, which, far from being opposed to relative simultaneity, is a
special case of it. The same is true of the relation of invariant succession. It is
to Robb's credit that he has constructed an axiomatic system whose only
supposedly undefinable concept is that of invariant succession, and that he has
thus shown that the four temporal relations of relative and invariant simultaneity,
and relative and invariant succession, can be defined in terms of a single relation.
According to the very nature of an axiomatic system, a primitive term cannot
be defmed in it without' a vicious circle. So, after according him the right to treat
the problem in axiomatic form, it would be wrong for us to ask Robb for an
explicit definition of 'conic order,' or what he means by the relation of before and
after in his restricted use of the term. But, before gmnting our consent, we shall
certainly ask about the empirical meaning of his fundamental relation. Admittedly
his explanation that event X took place before event Y, provided it could have
contributed to the production of Y, is too vague. What kind of possibility is meant
here? If anything which is not contradictory is eo ipso possible Oogical
possibility), then one might consider as possible the action of every event on
every other, since this implies no logical contradiction. If 'pOSsible' means
'compatible with the laws of physics,' we should point out that the only condition
imposed by the laws of physics is that the interval (in the Einsteinian
[Minkowski] sense) separating the two events must be negative - which
establishes only a symmetrical relation, inadequate for the explanation of
invariant succession. Moreover, this notion of possibility relative to the laws of
science could not be employed, without
94 THE CAUSAL THEORY OF TIME

circularity, in the definition of fundamental concepts which themselves appear in


the formulation of most of the laws. - Are we going to accept Kant's heroic
solution, which consists in saying that for one event to precede another (in the
restricted sense of invariant priority), the first must contribute effec-tively to the
production of the second? Schopenhauer's second objection warns us, however,
that one event can precede another without influencing it: if this objection did not
touch the Kantian defmition, it was because the principle of universal interaction,
adopted by Kant, presupposed the instan-taneous propagation of causal action -
an hypothesis explicitly rejected in Relativity. A more careful examination shows
that Schopenhauer's objection is not applicable, at least not directly, to the
definition of invariant priority in terms of the causal relation. This is because the
word 'event' in this definition denotes everything that occurs at a given instant in
a given place, a complete 'hic et nunc,' while the ordinary meaning of this term
is more general and allows us to speak of several events occurring in the same
place at the same instant: for example, the instantaneous values of gravitational
and electro-magnetic potential at a point in space, which form one and only one
'event' in the first sense of this term, constitute two distinct and coincident 'events'
in its second sense. Note that [if we use 'event' in its restricted sense] the
important notion of spatio-temporal coincidence boils down to tliat of logical
identity, in the first hypothesis, which clearly shows its inadequacy. We will
therefore use 'event,' which is becoming a fundamental category of empirical
knowledge, located at the juncture between relativity and the statistical ideas, in
its general sense, reserving the name of 'world-point' for the group of coincident
events forming a complete hic et nunc. The definition given above explained the
invariant priority of world-points in terms of the causal relation; applied to
events, this definition would take the following, different form: for event X to
precede event Y (in the restricted sense of invariant priority) it is necessary and
sufficient that an event coincident with X should have contributed to the
production of an event coincident with Y. Thus, according to this definition, one
event could (invariantly) precede another without contributing to its production
as Schopenhauer claimed. Yet this definition does not seem capable of providing
a satisfactory explanation of the fun-damental relation of Robb's system, since the
notion of spatio-temporal coincidence (i.e., of invariant simultaneity), which
appears in it along with that of the causal relation, is a temporal notion. In
supposing it to be in-definable, we would thereby give up the attempt at a causal
explanation of the temporal order.

We can arrive at a satisfactory causal definition of Robb's fundamental


THE AXIOMATIC SYSTEMS OF ROBB AND CARNAP 95

relation (and, more generally, of the four temporal relations) by' applying the
method of the causal decompositions of becoming. We have already shown in
connection with Lechalas how the relativity of time corresponds to the plurality
of causal decompositions. Under the hypothesis of absolute time, all events can
be arranged uniquely into classes so that each event is a member of one and only
one, class, and that given any two classes, one is always the cause of the other.
Under the relativistic hypothesis, becoming can be arranged into classes in many
different ways (we shall see later in what sense each arrange-ment corresponds to
a choice of reference system). Simultaneity and succes-sion are defmed relative
to a given decomposition; for example, two events are simultaneous relative to a
decomposition if they belong to the same class in this decomposition; the
relativity of simultaneity and succession therefore appears explicitly in the
definitions of these concepts. To defme invariant simultaneity and succession, we
have only to state that they take place in any possible decomposition. In this way
we obtain a causal definition of Robb's fundamental relation; for event X to
precede event Y (in the restricted sense of invariant priority) it is necessary and
sufficient that in every causal de-composition of becoming the class to which X
belongs be the cause of the class to which Y belongs. By putting this defmition in
place of the notion of priority in Robb's system, we can express all his
propositions in terms of the causal relation. Furthermore, the main result of this
system, that all of the topological and metrical properties of time, and even of
space, are strictly definable in terms of invariant succession, can be used as a
decisive argument in favor of the causal theory of time.

2. CARNAP'S SYSTEMS

This brilliant result was rediscovered in part by Camap, who expressed his ideas
on causal theory in a remarkable article in the Kantstudien, 56 as well as in a very
substantial form, though rather forbidding in appearance, in his sum-mary of
symbolic logiC.57 He claims to be able to show that the topological properties of
space are defmable in terms of the topology of time, and that time is defmable in
terms of causal action. He formulated this double thesis with great care and
precision, thereby facilitating our appreciation of its significance. Let us note in
passing that the first part of this thesis, the claim that the concept of space is only
derivative to that of time, seems charac-teristic of the recent phase of the causal
theory of time, since it is found, mutatis mutandis, in Robb and Reichenbach;
Lechalas, on the other hand, although very preoccupied with the "terrible
question of the nature of the
96 THE CAUSAL THEORY OF TIME

spatial relation," did not succeed in seeing its close relation to the causal
theory of time to which he subscribed. This fact is undoubtedly explained by
the influence of Relativity, which, while preserving the fundamental
distinction between time and space, established a much narrower connection
between them than was admitted before Einstein. His theory suggests that
the spatial and temporal orders of phenomena must have common roots -
which would explain the fact that having carefully analyzed the concept of
time, the causal theory seems thereby to have contributed to the clarification
of the concept of space. Moreover, the logical priority of time with respect to
space was already contained implicitly in Leibniz's doctrine, which defined
space as the order of simultaneous phenomena. Similarly, the parallel
treatment of the problems of space and time in the Transcendental Aesthetic
is contrasted with an asymmetry distinctly favoring the notion of time in the
Analytic.
Carnap has even shown that space can be reduced to time, and time to
causality, in three different ways. Thanks to logistic notation, he succeeded in
condensing his theory into a few dozen extremely precise axioms and
definitions: it is obvious that the resulting theory, so slight in appearance, is
the product of much fruitful labor. We shall try to summarize it without
having recourse to logistic notation. Let us note at the outset that Carnap only
undertakes to construct an axiomatic system of the topology of space-time;
by not attempting to solve the metrical problem, he derives results which are
also valid for General Relativity, while Robb's system, which furnishes a
homogeneous explanation of both the topology and the metric of space-time,
is valid only for Special Relativity. In his first system, Carnap shows that all
the topological properties of space-time are defmable in terms of two
fundamental and supposedly indefinable relations: the relation 'K' of spatio-
temporal coincidence and the relation 'Z' of priority in 'proper time.' It will be
remembered that in Special Relativity the 'proper time' of a particle is its duration
measured in a system of coordinates at rest with respect to the particle. The
twenty axioms and nine defmitions express the essential topological properties of
space-time in terms of the relations 'K' and 'Z'. Space-time consists of 'world-
points,' i.e., groups of coincident elementary events, each lasting infinitesimally
long and occupying an infmitesimal amount of space. These are defmed as
elements of the 'Held' of the fundamental relations. The world-points are grouped
into substantival lines, each of which constitutes the totality of events happening
to a given material or energetic element. There is substantial causal action
(Wirkungsbeziehung) of one world-point upon another if the first can be
connected to the second by a fmite number of substantival line segments joined
and placed end to end. This
THE AXIOMATIC SYSTEMS OF ROBB AND CARNAP 97

system differs from Robb's in the broader sense of 'event' which it employs,
which is precisely what necessitates the introduction of the concept of co-
incidence. In the second system, the only relation taken as indefinable is 'W',
substantival causal action. The following are some of its axioms: no event acts
upon itself; every event acts and is acted upon; if event X acts on event Y, there is
always another event Z acted upon by X and acting on Y ('density' of W). - The
third system is too technical to be summarized here; it does not differ essentially
from the first.
We will limit our criticism to an examination of the fundamental concept of
the second system, the most important from the point of view of the causal theory
of time. 'Substantival action', which plays the same role in this system as does
invariant priority in Robb's, is explicitly defmed in the first system. This
defmition lends a rather narrow sense to substantival action, restricting it to
action transmitted by successively coincident particles. Consequently,
Schopenhauer's second objection, as it relates to Robb's system, applies with
greater force to Carnap's, in which causal action in the general sense is replaced
by a very special causal relation. Furthermore, the choice of this special form of
the causal relation raises many more difficulties, of which we will enumerate
only the following two. Substantival action rests on the notion of proper time,
which becomes problematic in vacuo, as Carnap himself points out;
consequently, the temporal order of phenomena of spatial propagation is not
definable in terms of substantival action. Admittedly, this difficulty vanishes in
the corpuscular theory of light on the assumption that the ex-tremely problematic
localizability of Einsteinian photons was taken care of [cf. A. L. Akhiezer and V.
B. Berestetskii: Quantum Electro-Dynamics (1965)
- H. M. (1975)]: but then another arises, since substantival material or energetic
action becoming discontinuous in this theory, the temporal order of invariant
priority would not be defmed in it for all the events inside the cone terminating
with a given event, but only for those linked to the event at the vertex of the cone
by a chain composed of a finite number of lines joined end-to-end, and each
traversed by a material or luminous particle. Therefore substantival causality
cannot be used to defme the conic order in either the wave or the corpuscular
hypothesis. However, we will not insist on this point,
since the adaptation of the causal theory of time to quantum phenomena, and in
particular to the duality between waves and corpuscles, has not been attempted
until now; 58 it will be discussed only later, in the second part of this essay. What
seems to us more important from the point of view of the causal theory of time is
that by accepting the notion of substantival action as in-definable, one gives up
the attempt at an explanation of substantival duration,
98 THE CAUSAL THEORY OF TIME

which is an important component of the temporal order. Going back to


Figure 1 (p. 47), we can say that Carnap's causal theory allows us to decide
whether two events are on the same horizontal, or one is above the other, but
it assumes that we know when two points are on the same curve. However, we
shall see it propos of Russell that if we could answer the first two questions
independently of the third, the latter would then be susceptible of a causal
solution. This is the main reason why the choice of substantival action as the only
fundamental relation does not appear satisfactory to us. We believe that it would
be preferable to consider relation 'W', which plays the same role for Carnap as
invariant priority does for Robb, as being synonymous with the latter, and then to
complete Carnap's system with a causal defmition of priority in terms of causal
decompositions. This method would permit the reintroduction of the notion of
spatio-temporal coincidence into Carnap's system, where it is reducible to
identity,just as in Robb.
Let us note further that neither of these authors took the problem of
reversibility into account. This appears quite clearly in the very formulation
of the axioms. Thus, a glance at Robb's axioms suffices to convince us that
the replacement of the term 'before' by the term 'after' in each axiom and vice
versa, leaves them unchanged. But this only means that the axioms of the
system are inadequate for insuring an anisotropy of time, i.e., an intrinsic
difference between 'before' and 'after'. This is precisely what Robb's geo-metric
model of the conic order confirms. We have seen that to consider the family of
cones as an illustration of the temporal order, we somehow had to decide which
of the two cones emanating from a point was to be called 'upper' one. A choice is
necessary in order to interpret the system geometri-cally, but this choice is
arbitrary. The same is true of Carnap's axioms: if we replace the fundamental
relation 'W' with its converse, nothing is changed within the system. Must we not
conclude that the two systems would have gained much by considering, not the
anistropic and conventional temporal order, but rather the isotropic order,
independent of conventions on the temporal direction of becoming?

3. THE EPISTEMOLOGICAL PRIORITY OF THE CAUSAL RELATION

Here is a question of a more general order: how shall' we evaluate the epis-
temological significance of these deductive systems, which resolve physical
becoming into causal chains formed of world-points? What part does intuition
play in these conceptual constructions, which are as coherent as they are
removed from what is immediately given? We do not perceive elementary
THE AXIOMATIC SYSTEMS OF ROBB AND CARNAP 99

events linked by causal action, but data of fmite dimensions -' colored sur-faces
which change their shape and move, prolonged sounds, etc, From this universal
becoming, common sense carves out processes delimited in time and space, such
as the flight of a bird, the fall of a leaf, the whistle of a train, This sorting process,
indispensable in practice, is so familiar that it requires considerable mental effort
not to do it, or to do it differently, Science pushes this common sense activity to
the extreme, arriving at processes more and more narrowly bounded in space and
time. At its limit, which, practically speaking, cannot be reached, this dissection
would result in processes occupy-ing an infmitesimal amount of space and lasting
an infinitesimal amount of time; that is, in the elementary event, the world-point.
However, this passage to the limit, and this bold extrapolation of a practically
bounded process, arise from the very nature of science, which aims at absolute
objectivity, the reduction of the observer's rale to the minimum possible and his
replacement by instruments and calculation. 59 But the analysis of becoming
always implies a subjective element, since it is relative to both an observer and
the choice of a system of coordinates which permits the observer to perceive the
processes isolated in becoming all at once, or to submit them to calculation.
However, as the sorting out advances, as the pieces themselves are divided into
more pieces, etc., the subjective factor becomes less and less Significant; in
pushing the division to the limit, that is, to the world-point, we would eliminate
the subjective factor entirely. The simplified physical universe at which we would
arrive would be composed of world-points enjoying an abso-lute order,
independent of every system of coordinates and every convention of
measurement. This is precisely the topological order of neighborhood envisaged
by Carnap, the order of coincidence of General Relativity.

Thus, the notions assumed to be indefinable in the systems of Robb and


Carnap are the results of very complex mental processes far removed from
what is intuitively given. We have no intention of doubting the logical worth
of these authors' efforts at defining the properties of space-time with a
minimum of supposedly indefmable notions. Notice however, that the role of
the definition of the fundamental concepts of science can be viewed quite
differently: instead of regarding all that science teaches about a fragment of
reality as something ready made, and instead of being satisfied thereby with
every defmition which attributes to the concepts analyzed only those prop-
erties which arise from these teachings, we might also require that the de-
fmition somehow reflect the course of scientific research itself, and that it
take into account not only the raw result, but also the method which led to it.
Thus, with respect to the causal defmition of time, if we limit ourselves to
100 THE CAUSAL THEORY OF TIME

an axiomatic analysis of what science teaches us about time and causality, the
question of knowing whether or not the experimental determination of an
arbitrary causal relation presupposes the notion of time does not arise: it is not
the connection between our knowledge of time and our knowledge of causality
which comes into play, but rather the relation between these concepts themselves.
It is nevertheless undeniable that definitions of the second type have a quite
different importance from those ofthe first, which, although taking into account
the intrinsic connections among the ideas, give a distorted view of the cognitive
process.
This seems to be inherent in the very nature of axiomatic analysis. Axio-
matizing empirical knowledge of time essentially amounts to considering this
knowledge as based on a certain set of propositions about time which are
supposed to be true. Obviously this set contains not only those propositions
which, for some reason or other, we may wish to formulate explicitly, but
also all propositions obtainable from these by pure logic. This closed 60 set of
propositions (Le., it cannot be enlarged by any operation of pure logic)
expresses our knowledge about time from the point of view of axiomatic
analysis. A closed set necessarily contains an infinite number of propositions, but
there are infmitely many ways of choosing from it some finite group of
propositions from which all the propositions of the set can each be deduced by
the application of a fmite number of logical operations. Such a group is called an
axiomatic basis of the set, and its choice, an axiomatization. Therefore an
axiomatization necessarily idealizes the knowledge of time by substituting an
infmite set of propositions for the fmite set actually stated. It artificially simplifes
the problem by substituting a closed set of propositions for a knowledge which is
constantly growing.
These remarks are not meant to minimize the intrinsic value of the axio-matic
method which Carnap and Robb have applied to the causal theory of time and
which has become so important in contemporary philosophy of science. 61 Its
partisans have clearly been inspired by the axiomatic method of geometry, which
they have tried to use in a new domain, that of the epistemological foundations of
science. This generalization of the axiomatic method is undoubtedly a major
achievement in the recent development of the theory of knowledge. Nevertheless,
let us not forget the dangers which this new method carries with it. Basically, the
method roughly amounts to this: 'Tell me what my fundamental relation is and I
will tell you the rest' . Or, in the particular case under consideration: 'Tell me
what causality is, and I will tell you what time, and even space, is'. What would
happen if, in order to explain the fundamental relation, it were necessary to have
already explained
THE AXIOMATIC SYSTEMS OF ROBB AND CARNAP 101

the 'rest' (or even a part of the 'rest')? Would it necessarily follow that the theory
is false? Obviously not, since the axiomatic theory maintains that to understand
the 'rest', it is enough to understand the primitive relation, which does not deny
that understanding the 'rest' might be necessary for under-standing the
fundamental relation; but anyone attempting to explain the 'rest' in terms of the
primitive notion would be involved in a vicious circle. The proof, attempted by
Reichenbach, according to which the fundamen~al relation of causality does not
presuppose the temporal order, does not employ the axiomatic method.

4. THE ONTOLOGICAL PRIORITY OF THE CAUSAL RELATION

The fact that the temporal order is definable in terms of the causal order tells us
nothing about their cognitive relation and does not involve the epistemolo-gical
priority of the causal over the temporal. Can we conclude from this, with some
representatives of the causal theory of time, that there is a real priority of the
causal order, which must be considered as a more fundamental feature of
becoming than the temporal order? Can we build the causal theory of time into a
metaphysics which asserts that the universe is the field of a certain relation W,
called causal action, time and space being merely some of its structural
peculiarities? This metaphysics is not inevitable for the following two reasons:
first of all, one can argue that it is incorrect to give science any metaphysical
interpretation whatsoever - perhaps science admits only of a pragmatic
interpretation, its formulas and theorems constituting a set of practical rules for
orienting us in the world of phenomena as well as fore-seeing its development.
Secondly, even if we admitted the possibility of a metaphysical interpretation of
the causal theory of time, such an interpre-tation would merely be compatible
with science, and would not necessarily follow from it. To illustrate this point,
and also to determine what is implied by the fact that one physical notion is
definable in terms of another, let us take a well-known geometrical example: we
know that all of Euclidean geometry can be constructed by taking as undefmed
only the relation of four spheres having a point in common. In this system, the
relation R is the only primitive notion, since the spheres can be defined as its
field. Hence, taking relation R as our starting point, we will defme the
fundamental concepts of geometry, point, line, plane, etc., whose essential
properties will be expressed in terms of the axiomatic or derived properties of
relation R. Must we con-clude from this that space is only a certain order of the
spheres, and that a point in particular is only a certain group of spheres?
Obviously not, since, as
102 THE CAUSAL THEORY OF TIME

is done in elementary geometry, we can just as well take the concept of point as
indefinable and define a sphere as a class of points equidistant from a fixed point,
while before, a point was defined as a class P of spheres having the following
two properties: (1) any four elements of P are linked by the relation R; (2) no
element can be added to P without depriving it of the property (1). Thus, in the
first system the point is a class of spheres, and in the second, the sphere is a class
of points. If both theories were true and stated in the same language, by
combining them we would obtain the result that a point is a class of classes 0°
points and that a sphere is a class of classes of spheres, which is absurd, since a'
class, being of a higher logical type than that of its elements, is never identical to
them, the relation of identity holding only between objects of the same type (a
fortiori), a class of classes of element X cannot be identical to element X.
Obviously, one must ask why the partisans of the two systems use the same
names for objects having such different logical structures. Two answers come to
mind: (1) the same names are used because the two systems are supposed to
correspond to an objective reality, and their homonymous symbols correspond to
the same fragments of that reality; (2) the identity of the symbols does not entail
any reference to an objective reality transcending the systems, but only expresses
the existence of a certain correspondence between the two systems, whereby
homonymous symbols play analogous roles. (We would immediately discard the
first answer if we took geometry to be a free construction of the mind, aiming
only at internal consistency.)

What is important for us is the interpretation of geometry as a basic science


of nature, smce the supporters of the causal theory of time also claim to be
putting forth a theory of time applicable to physical reality. But, for reasons of
economy, we prefer the second answer (assuming it to be compatible with the
analysis of the two systems in question), especially since it dispenses with the
strange hypothesis that the same fragment of reality could correspond to quite
different mental symbols. To show more clearly that a concept definable in one
particular axiomatic system but indefinable in another must be inter-preted in
this way, let us point out that we can consider any deductive theory as attributing
a certain number of properties, enunciated in the axioms to its indefinable
notions: thus, Hilbert's axioms characterize the class of points, straight lines,
planes, and the relations of congruence and position (,between-ness') in a
particular way. On the other hand, all the axioms of the system can always be
replaced by a single axiom, defined as their 'logical product'. This unique axiom
would then define the properties of the fundamental entities of the system in
terms of pure logic. We can therefore put the entire logical system
THE AXIOMATIC SYSTEMS OF ROBB AND CARNAP 103

into the symbolic formF(xf,xi, ... ,xL ... ,Xh,X~ .. . x~), 'x~'designat ing the m-
th symbol of the logical type of order n admitted in the theory (we assume
that these types are arranged in an arbitrary way into an infinite series while
'F' designates a propositional function definable in terms of pure logic). Thus,
in Hilbert's system we would write: 'F (class of points, class of straight
lines, ... , relation of congruence, ... )'. The function F is ['typically ambi-
guous' and] 'systematically equivocal', Le., its structure prescribes only the
relations among the logical types of the variables, leaving open the choice of
the type of minimum order. We will say that two systems Sand S', with the
unique axioms: F (x~ ... x;), G(y~ .. . Y~'), respectively, are equivalent
if in S one can define a group of concepts ji 1 ,ji~ .,. such that we have
)d'f' S' dfi f
onecan

-{;;1 -2 1 -1-2
G
V1'Y1 ... ,a~ 1 m emeagroupo conceptsx1 ,x.1'"

~ch that we have F(i!, x: ... ), the functions F and G differing from F and
G only in logical type. This is precisely the case with the two systems we just
discussed. It is obvious that the equivalence of the two systems depends only on
the logical structure of functions F and G, which constitute the only data of the
problem, and that no relation to a reality external to the two systems enters into
the definition of equivalence. We conclude that the possibility of defming the
entities of one system in terms of the entities of another does not depend on the
nature of these entities but on the structure of the axiomatic systems which
correspond to them. The same is true of the causal defmitions of time. If our
analysis is correct, these definitions do not show that time is a mode of causality,
but only that our axioms of time and causality constitute two equivalent
axiomatizations of the same closed set of propositions. If the principle of physical
determinism appears as a definition in one of these systems (as seems to result
from the causal theory of time), it could appear as an axiom or even a theorem in
the other ,just as the elementary properties defining the circle in some axiomatic
system of geometry can be deduced in some other system from its isoperimetric
properties. Thus, Jordan, in the article cited above, considers this principle as
resulting from the hyperbolic form of the partial differential equations of
theoretical physics.
It is precisely this great flexibility of the axiomatic method, whose use leaves
open the choice of not only the indefinable concepts and undemon-strable
propositions [axioms] but also the 'logical type' of the variables employed, which
accounts for the invariance of the theories with respect to the epistemological
transformations we considered in connection with Lechalas: this is because the
epistemological point of view presupposes a definite choice of concepts,
propositions, and types, a choice which the axiomatic method facilitates by
specifying on what it must bear, but which it
104 THE CAUSAL THEORY OF TIME

does not determine. Moreover, it may be that in certain cases this choice is
impossible or superfluous, for example, we might consider phenomenalism
and realism as two equivalent axiomatizations of empirical knowledge even
from an epistemological point of view. But this again would not be a con-
sequence of the axiomatic method itself. Similarly, the question of whether
the causal definition of the temporal order is the only one which is satisfac-
tory from an epistemological point of view (in that it gives scientific
positions an experimental sense with respect to time, and makes them
veriftable), will be answerable only by making a careful analysis of the
sources and multiple forms of empirical knowledge of time.
v

THE RELATIVISTIC PHASE OF THE CAUSAL


THEORY OF TIME: THE WORK OF REICHENBACH

1. THE CAUSAL THEROY

Reichenbach, who has perhaps contributed the most to the recent develop-
ment of the causal theory of time, approached his axiomatic research from
the point of view of epistemology, that is, the theory of physical knowledge.
The Axiomatik der Relativistischen Raum-Zeit-Lehre (1924) [translated
as Axiomatization of the Theory of Relativity (1969)] , his great attempt at a
philosophical interpretation of Relativity, a book which is essential to an
understanding of the logical structure of Relativity, is almost entirely domi-nated
by the idea of causal action of which space and time are only expres-sions of
structural features. Thus, in Special Relativity he deduces the spatial order from
certain properties of the temporal order, itself based on the fundamental
properties of causality, and tries to prove that within the very general systems of
coordinates admissible in Einstein's Theory of Gravitation these fundamental
properties are preserved, along with the temporal order which they engender. So,
for Reichenbach, the causal theory of time (and space) constitutes the
philosophical import of the whole of Relativity.
We will approach the statement of the causal theory as it is found in three
of Reichenbach's writings,62 in terms of a method which seems to have been
the point of departure for his studies and which we might call the 'method of
messages'.63 This method systematically develops the well-known
Einsteinian distinction between the two kinds of simultaneity - the
simultaneity of events occurring in the same place (the fundamental relation
of 'coincidence'), and that of events occurring at different points in space,
defined by the coincidence of two light rays whose departures coincide with
these two events respectively at a point equidistant from the two points of
space. Thus,64 the Einsteinian defmition explains a certain temporal relation
in terms of another more fundamental relation and a causal process.
Reichenbach goes on to generalize this procedure by introducing a causal
process into the definition not only of the simultaneity of spatially separated
events, but also of the succession of any events whatsoever. One of the main
results of his work was to divest Einstein's definition of this arbitrary and
somewhat artificial aspect for which he has so often been criticized.
105
106 THE CAUSAL THEORY OF TIME

Here is how the 'method of messages' solved the problem of the genesis
of the temporal order. The universe is made up of events. Among these
events, there are some which all happen to a single material point (this is
essentially
a definition of the material point in terms of the symmetrical relation 'hap-pening to
the same point', the vicious circle being only apparent): these constitute the world-
line of this point. An observer, able to follow all the events succeeding one another at
point P, should know that all of these events take place at the same point P, but he is
incapable of deciding intuitively which of two events, El or E 2 , both taking place at
P, is earlier - or, if he thinks that he knows it intuitively, he is called upon to renounce
his intuitive knowledge of succession in this case. To resolve the question in
conformity
with the causal theory, he will send a messenger to a suitably chosen neigh-
boring point, Q (this point will playa special role in what follows); if the return
of the messenger, who will have left P simultaneously with E 1 (notice that this is
simultaneity at the same place), coincides with E2 at pointP, we will say that E 2
is later than E 1 ; if, on the other hand, it is possible to send a messenger from P
so that his departure coincidences with E 2 and his return with E 1, we will say
that E 1 is later than E 2 • For events sufficiently separated in time, any messenger
could be used, but to determine the succession of two events close to each other,
the messengers must be as rapid as possible:
therefore, in this case we will prefer to replace the living messenger with a
light ray65 which will leave from P, be reflected in a mirror at Q, and return
to P; any event coinciding with the departure of the beam will be said to be
earlier than an event at P coinciding with its return. Obviously, one of the
axioms will be that if two messengers are sent from the same point it is
impossible that both the departure of messenger 1 coincides with the return
of messenger2, and the departure of messenger2 also coincides with the
return of messenger 1 (the impossibility of closed causal chains).
At the point, two difficulties arise. First of all, how will the observer at P
be able to tell that it is the same messenger, and in particular, the same light
ray which coincides with P twice in succession? Secondly, even granting that
he has a way of identifying the messenger, how will the observer be able to
distinguish between departure and return, since he has just been called upon
to 'renounce' his intuitive knowledge of succession? To my knowledge
Reichenbach did not make known his answer to the first question (to which
we will return in connection with Russell's theory), but he did give a very
ingenious - and very debatable - answer to the second. He assumes that a
message coinciding with Q can be slightly changed, the change bearing witness
to its passage at Q while still safeguarding the identity of the process. If the
THE WORK OF REICHENBACH 107

message is a light ray, one could change its intensity, frequency, polarization,
etc.; if it is a material body which goes from P to Q and back again, one could
mark it in some way or other at Q, with a piece of chalk, for example. Suppose
then that the messenger was 'marked' in a certain way when it passed to Q and
that the observer at P had noticed this mark on only one of its two trips to P: we
will then agree to consider the coincidence a/the marked messenger with P
as later than that of the unmarked messenger, and the two coinciden~es will be
distinguished as the messenger's departure from and return to point P.
So far, then, the method of messages permits us to establish the succession of
two events at the same point in space. To extend it to the general case of a
succession at arbitrary points in space, Reichenbach formulates a general
proposition, which he calls 'the principle of marking' (Kennzeichenprinzip)
which is basically nothing but a new definition of the causal relation. 66 This
definition, essential for Reichenbach's causal theory since it permits us to
distinguish cause from effect without examining their temporal relation (and
consequently to define succession in terms of causality), can be summarized
approximately as follows: a pair of events, El and E2 , are causally related if they
are found together in every region of the universe (in Minkowski's sense). If in
this pair, small variations in E I ('marks' applied to E I) are always accompanied
by small variations in E2 , and not vice versa, then EI will be taken as the cause
and E2 as the effect. For example, in the case of the free fall of a heavy body: E I
will be coincidence of the stone with a point in the air, E 2, the coincidence of the
same stone with a point on the earth's surface. If I mark the stone with chalk,
thereby slightly changing event E I, I will notice that E2 will also be changed (the
stone will still be marked); but if I applied the mark to E2 , EI would remain
unchanged. I would then conclude that E I is the cause of E 2, and consequently
prior to it.
It is easily seen how the marking principle permits the extension of the causal
defmition of succession to events occurring at different points in space: one has
only to replace the 'return' of the messenger with its 'arrival': event
E I is earlier than event E 2 if there is a messenger whose departure and arrival
coincide with EI and E2 respectively. To distinguish the 'departure' of the
messenger from its 'arrival' we will use the marking principle. For example, we
will make different marks on the messenger when it passes to points PI and P2
(where E1 and E2 take place): if the mark applied at the time of the passage to PI
is also observed by the observer located at P2 , while a mark made by the
observer at P2 is not found by the observer at PI, we will say that the messenger
left PI and arrived at P2 . The opposite will take place if it is the mark made at P 2
which is transmitted by the messenger.
108 THE CAUSAL THEORY OF TIME

Notice that no notion of measurement enters into this defmition of before


and after. To use the language of geometers, who call 'topological' those
properties of spatial figures and of space itself which do not depend on mea-
surement (or, technically speaking, are invariant with respect to continuous
one-to-one transformations), we can say that the topological properties of
succession are defmable in terms of the causal relation. Is the same true of
simultaneity? It has sometimes been maintained that simultaneity is a purely
metrical notion, since it expresses the equality of two numbers having re-
ference to time. Reichenbach correctly points out that, like succession,
simultaneity also has a topological aspect, even if we continue to define it as
the equality of two numbers indicating time. Let us suppose that we have
established a one-to-one correspondence between the sequence of real num-
bers and all the events succeeding one another at a point in space, so that the
earlier of two events corresponds to a smaller number. We have done this by
using a suitable device, namely a clock, placed at this point (this is a
topological definition of the clock, since it applies to normal clocks as well as
to those which are as 'irregular' as may be desired). By sending messengers
from this point into all the regions of the universe, we can use the readings of
the central clock to regulate the others and thus establish a universal
simultaneity. It is clear that the causal theory of time imposes only one
restriction on the regulating process: it is that if event E 1 takes place at point
PI ,and is the cause of E 2 at P 2, then E 1 must precede E 2, i.e., E 1 must
correspond into a smaller number. Suppose that a messenger, having left Plat
instant t 1 (which means that the messenger's departure coincides with an
event to which the topological clock at PI assigns the number td, travels to
P2 simultaneously with the occurrence of event E and then returns to PI at t2 ;
what number must we assign to E in order to preserve the causal postulate?
Evidently, it suffices that t ' = tl + €(t2 - td, the number € satisfying the
inequality of 0 < € < 1. We obtain Einstein's first definition, in terms of light
rays, by making € = ~. From the point of view of the causal theory this
determination is absolutely conventional (although not arbitrary: there are
good and bad conventions). Any other metrical defmition of time satisfying
the condition imposed on the number € by the above inequality is equally
admissible from the point of view of the causal theory of time. We can even
make € vary in space, as Reichenbach has shown in two curious examples
(which is moreover perfectly admissible in General Relativity). Thus,
according to Reichenbach, the temporal order has two essentially different
components from the epistemological point of view (although not
distinguished by the physicist): the topological component, expressing a
THE WORK OF REICHENBACH 109

causal structure inherent in becoming, and the metrical component, conven-


tional and relative to the needs of science. Any passage from the topological
order to the metrical order implies conventions of measurement.
If the convention related to simultaneity did not figure in traditional physics,
it is because physics rested on another topological conception of time, and, in
particular, of simultaneity.· The causal theory of time provides an excellent
explanation of the transition from the classical conception to the relativistic one
whose paradoxical aspect it clears away, provided one takes the principle of action
by contact seriously. Let us consider two bodies separated from each other: we
will say (in the causal theory) that state £1 of body C1 is earlier than the state £2
of body C2 , if a physical action stemming from £1 was able to contribute to the
production of £2; if, on the contrary, an action caused by £2 had contributed to the
production of £1, then £1 would be later than £2. For them to be simultaneous, it is
necessary and sufficient that they cannot act on each other. Up to this point, there
is nothing to distinguish Einstein's time from classical duration. But add that any
physical action is transmitted by contact with a fmite velocity: any message will
take some time to go from C 1 to C2 and back again. In order for event £ which
will take place within C1 after the departure and before the return of the
messenger to be able to act upon an event in C 2 which coincides with the passage
of the messenger at that point, this action obviously must be propagated with a
velocity greater than that of the messenger; on the other hand, if at the instant of
the messenger's passage, an action must be performed by C 2 on an event situated
in C1 , this action should be propagated with a velocity still greater than that of the
messenger. But if the message is trans-mitted with a maximal velocity,
unsurpassable by any physical process, no physical action will be possible
between all the events taking place in C 1 after the departure and before the return
of the messenger on the one hand, and the instant of the messenger's arrival in C 2
on the other. Consequently, in conformity with the causal theory of simultaneity,
this instant in C2 will have to be considered as simultaneous with all the instants
contained in the time interval separating the departure and return of the messenger
in C1 • In other words, under the hypothesis of the existence of a maximal velocity
of prop-agation, there corresponds to the state £2 of body C2 an entire continuous
series of states of body C 1 such that no causal chain can link £2 with any element
of this series. All the elements of the series will be considered simul-taneous with
£2, and only by convention could we select one rather than another. Moreover,
there will be every advantage in chOOSing the Einsteinian convention, with
regard to a frame of reference. Notice that to explain the
110 THE CAUSAL THEORY OF TIME

relativity of simultaneity, it is not sufficient to affirm that all action is trans-


mitted with a finite velocity; we must add that there exists an action with a
maximal velocity of propagation.
As Reichenbach points out, in the absence of a preferential velocity, we
can still define an absolute simultaneity by approaching the limit through
faster and faster messages. Here again the quarrel between the partisans of
absolute and relative time bears on a problem of causality. In speaking of a
maximal velocity we do not object to the presupposition of the notion of
time. The existence of a maximal velocity really expresses a topological
property of light, definable in terms of the causal relation. It could be stated
approximately as follows: two events which coincide with the departure and
arrival respectively of an arbitrary messenger can always be linked by a light
message (if necessary, the path of the light ray can be lengthened by the use
of mirrors).
The relativistic contention that the speed of light is unsurpassable by any
physical process has often been the object of more or less ingenious criticism. We
will concentrate on one of these, since Reichenbach drew an interesting argument
for the causal theory of time from its refutation. By using precisely the
elementary properties of light, this criticism claims to prove the existence of
processes taking place with velocities as large as may be desired. For ex-ample,
take a lamp emitting a very narrow beam of light through a slit and fixed on a
base which rotates uniformly about a vertical axis. This beam will successively
illuminate the points of a horizontal circle of arbitrary diameter traced about the
lamp. Let n be the number of turns effected by the lamp in a second, d, the length
of the diameter, and c, the speed of light. It is imme-diately seen that the
illuminated point will be displaced on the circle with a velocity greater than c,
provided the number 1T • n • d satisfies [1T • n • d> ~ ]. By tracing larger and larger
circles one obtains greater and greater velocities which can be used to defme an
absolute simultaneity and thus refute the relativistic thesis. However, a more
attentive examination of the displacement in question dispels the illusion. The
calculation of the velocity of this dis-placement presupposes the uniform and
isotropic propagation of light, which, in turn, as we know, presupposes the
Einsteinian defmition of simultaneity. It will suffice to change the definition of
simultaneity (by locally varying e, equal to ~ for Einstein) in order to change the
temporal order of events on the circle entirely. Thus, the process in question does not
have an intrinsic temporal order, which depends on a conventional simultaneity. But it
is evident that this displacement is not a causal process either: if we place an opaque
body between the lamp and the circle, the luminous point will
THE WORK OF REICHENBACH 111

disappear from the arc which lies in the shade, while the rest of the processes
will be unchanged, proving that the consecutive states of the process do not
determine one another. Thus a non-<:ausal process, an 'unreal sequence' ,67 as
Reichenbach calls it, has no intrinsic temporal order. These two properties
explain each other and reveal the unreal character of processes occurring
with a velocity greater than that of light. This also explains the fact that
unreal sequences, which always presuppose a conventional simultaneity,
cannot defme an absolute simultaneity: starting from different conventions
we would arrive at simultaneities which are 'absolute', but which are
nevertheless incompatible with one another.

2. CRITICISM OF THE PRINCIPLE OF MARKING

In criticizing these interesting ideas of Reichenbach, we will limit ourselves to


his principal assertions concerning the conventional character of relative
simultaneity and the role of sensibility in the knowledge of time, beginning with
an examination of the principle of marking. Perhaps its initial convincing
appearance is due only to a linguistic confusion. Let us return once more to the
example of the stone which falls vertically under the influence of its own weight.
The two events: (1) the coincidence of the stone with a portion of the air, and (2)
its coincidence with a point on the surface of the earth are undoubtedly linked by
a causal relation. Which of the two precedes the other in time? To apply the mark
method we will effect a slight change in each event, too small to affect the
process, but capable of being transmitted from place to place - for example, we
will mark it with red chalk simultaneously with its upper position, and with white
chalk with its lower position. We will then note that the red mark will also be
found when the body is in its lower position, while the white mark will not have
appeared in the upper position. Thus, in keeping with the mark method, the upper
position is prior to the lower one. This reasoning, even if it were correct, is no
more than an enthy-meme. In fact, the mark made at the beginning of the process
will be found at its end only if we have taken care that it should not be erased
during the fall. In other words, for a signal to be transmitted along a causal
process (in the temporal direction of past-future) certain 'boundary conditions' (cf.
Jordan's statement in connection with Lechalas, cited above) must be fulfilled
during the interval of time separating the two events linked by the causal relation
(and to be certain that these conditions were fulfilled, we would have to know
which events take place in this interval, i.e., we would have to be able to locate
the events in time with respect to one another). As for the white
112 THE CAUSAL THEORY OF TIME

mark, applied at the end of the process, everything depends on the meaning
of 'to mark'. If 'to mark an object at a given instant' means 'to apply a mark to
it which it did not have before', the mark method is reduced to a tautology
and tells us nothing about the temporal order of events, since, in order to
know if the stone has been marked at a given instant, it is not enough to
observe the sign at that instant, but we must also be sure that the stone did
not have the mark before; we would have to be able to distinguish before
from after, and for this, the method would be of no use. If, on the other hand,
'to mark the stone at a given instant', is to make sure that the sign appears on
it at that instant, without regard for what happened before, we will not be
able to conclude from the appearance of the sign in the stone's final position
that it was not there in the initial position. The two positions of the stone play
exactly symmetrical roles with respect to marking, which therefore cannot be
used to distinguish before from after.
Here is another difficulty. The mark method seems to attribute an irrever-sible
character to becoming. Why then did the author support it with examples from
mechanics and optics, whose phenomena certainly appear to be rever-sible? In
his first book, Reichenbach did indeed remark that the irreversibility of the
phenomenon used for marking could be essential for the validity of the principle,
but he did not go on to settle the question. To clarify this point, let us imagine a
room with perfectly elastic surfaces in which an elastic ball falls vertically from
the ceiling to the floor; obviously it will bounce back to the ceiling and this
process will repeat itself indefinitely. A man in the same room who wanted to
distinguish before from after in the successive coincidences of the ball with the
surfaces of the room could apply the method of marking by covering the floor
with a thin coat of paint: will he then be justified in concluding that all the
coincidences of the ball marked by a spot of paint are later than that one when
the ball was unmarked? Suppose that the ball has executed two movements: from
ceiling to floor (unmarked) and from floor to ceiling (with a mark acquired upon
contact with the floor): there will then be two coincidences with the ceiling, one
when the ball is marked. This is certainly a typical case of the mark method, yet
we believe that the observer, having decided to 'renounce' his intuitive
knowledge of time, will not be able to decide which of the two coincidences
precedes the other without having recourse to irreversibility. Under one
hypothesis he will say that the marked ball fell to the floor with a uniform
acceleration, lost its mark upon making contact, bounced back up unmarked;
somehow, the floor would then have absorbed the mark. Under the other
hypothesis, the ball, having fallen un-marked, would have acquired the spot upon
contact with the floor and kept
THE WORK OF REICHENBACH 113

it on its trip upwards to the ceiling. Both hypotheses are possible since both
verify the same laws of physics, given different initial conditions. The only
debatable point would be the double effect of the contact with the floor, with
respect to the relative motions of the bodies in question. But it seems clear
that if the phenomenon of the acquisition of the spot is reversible, the two
hypotheses will be equally possible for our observer; if not, it will not be the
mark method which will have informed him of the temporal order of the
observed events, but this irreversibility. Therefore, we are led to the same
conclusion as that a propos of Lechalas' theory: in a reversible universe,
succession cannot be defined in terms of the causal relation; the definition
based on irreversibility is not a causal definition; rather, the statistical de-
fmition of succession must still be combined with the causal defmition of
simultaneity. The mark method does not change this conclusion.

3. CRITICISM CONTINUED: SUBJECTIVE AND OBJECTIVE TIME

Let us now pass to an examination of the 'renunciation' of our intuitive


knowledge of time to which Reichenbach invites us. Evidently, Reichenbach
speaks here as a physicist, which in our time often implies a profound dis-
respect for common sense; he makes a point of renouncing the cognitive
resources of common sense and yet claims to know even more than it does
about the sensible world. What, in effect, is the typical behavior of the
physicist, as seen by the layman? He is seen ordering apparatus from various
factories and then arranging it in a particular way. He sets his machines to
work one after another, anxiously watching to make sure that everything is
going according to a prescribed plan. At a certain moment he declares that he
has seen enough, jots down the final positions of the needles, takes his notes
and starts calculating (sometimes for many months); finally numbers appear
which make the physicist's heart beat faster: his theory is verified.
To the uninformed spectator, the physicist does not seem to have left the
world of common sense for a single instant, nor renounced his intuitive
knowledge of it. His devices are 'things,' more or less strange or complicated, but
nevertheless of the same sort as the familiar objects of daily experience. The
handling of both seems, in principle, to require the same aptitudes. His
observations bear on the positions and displacement of material objects, as do
everyone's observations. His calculations are carried out by ordinary means on
pieces of paper, as are those of everyone else. Yet, one thing remains strange:
when asked what he has just accomplished, the physicist tells a tale which does
not seem to be of this world, a tale in which the author has
114 THE CAUSAL THEORY OF TIME

nevertheless managed to accomplish real miracles. If a needle moves, for


him it is the intensity of an electric current which has varied; if a flash of
light appears, it is an electron that has changed its trajectory; etc. Yet, the
same physicist may deduce the radius or duration of the universe from his
calculations, and will make himself heard on the other side of the earth,
something that the loudest shouter, as an exclusive supporter of common
sense, would never have achieved. Doesn't he bring to mind the
somnambulist who accomplished inconceivable feats, all while sleeping?
And does not the physicist's marvelous success inspire the same confident
terror as does that of the somnambulist? And is there not a nascent suspicion
that this mysterious world where the physicist claims to live during his
research and from which he seems to draw such efficacious insight about our
own universe, is a domain of dreams and illusions, suggested to the physicist
by his extra-ordinary capacities, by his legendary dominion over our world,
an empire so marvelous that it does not seem possible that he himself should
be of this world?
The physicist himself, asked about the strange ambiguity of his attitude, might
reply in this manner: "From the sensory data furnished by experience, science
aims at constructing a mathematical universe, more precise, more schematic, and
also less fragmentary than the universe of common sense. The purpose of this
construction is to be able to describe these data with a maxi-mum of economy
and to predict with a maximum of certainty and precision. Thus science imitates
and extends the work of common sense, which itself, guided by instinct and
habit, replaces the chaos of the immediately given with a more ordered universe.
Science and common sense use, in principle, the same data of sensible
experience, yet, thanks to different principles of con-struction, they arrive at
different results. But there are also differences in the very choice of data. In
principle, common sense uses all sensory data; to it hot and cold, dry and humid,
light and heavy, seem to be categories just as important as before and after, inside
and outside. As science evolves, an increasingly rigorous choice of utilization
data is imposed. The physics of heat owes its creation to the thermometer, which
permits the replacement of sense data of temperature with visual data. Thus it has
been possible not only to increase enormously the precision of the measurement
of temperatures in the normal range, but also to study very low and very high
temperatures, which otherwise would have remained inaccessible. It is in this way
that at different times in the course of scientific evolution we have renounced the
use of immediate data corresponding to velocity, mass, electricity, etc. Sometimes
we do not entirely discard a datum, but we limit its use by subordinating it to
THE WORK OF REICHENBACH 115

another which has become more fundamental. This is how coincidence tends
to become the fundamental datum in the spatial order.
"Now the causal theory of time, in its recent form, does seem to imply an
epistemological interpretation of relativity. In the construction of the
temporal order of the physical universe, the scientist must abandon the use of
the intuitively perceived succession of phenomena as a fundamental datum,
replacing it with a study of causal relations. Of course, he does not abandon
it in practice, just as the experimenter, in spite of all the progress of
thermometry, makes constant use of his sense of temperature. But the funda-
mental datum in the construction of objective duration will be the ascertain-
ment of a causal relation, and not a perceived succession. As a physicist, this
is how I understand Reichenbach's all too concise statements."
Thus might the physicist speak who was a supporter of Reichenbach.
Would he succeed in convincing a partisan of the intuitionist theory of time?
We think not. If the latter were to take the floor in his turn, he would no
doubt reply to the physicist-philosopher in this way:
"You say that to decide which of the two phenomena, A or B, precedes the
other in time, it is sufficient to determine which causes the other. Sup-posing it
were sufficient, how would you be able to decide that a given event was the cause
of another. You would undoubtedly take a little trip through the universe to make
sure that everywhere A is followed by B. Consequently, to know that the first
engendered the second, it would be necessary to know their relation of
succession, which would be a vicious circle from the point of view of your causal
theory of time. No doubt you would try to remedy this by trying to give an extra-
temporal defmition of the causal relation. I would very willingly follow you in
this task, for it would always be worthwhile, even for someone who was not a
supporter of the causal theory of time, to be able to analyze causality in purely
logical terms, thus taking another step toward the rationalization of our
conceptual apparatus. However, I do not see how we go about defining causality
independently of time, preserving its former meaning, or at least its former logical
extension. Would you start by abandon-ing the distinction between cause and
effect, trying only to pick out 'pairs of events linked by the causal relation' where
the roles of cause and effect have not yet been assigned? But how would you
recognize these pretemporal pairs? It is not enough to say, as is usually done, that
one element of the pair is always 'accompanied' by the other, unless you
understand this accompani-ment in the sense of a spatial proximity, which itself
undoubtedly presupposes in turn the temporal notion of simultaneity. 'Two events,
£1 and £2, which always seem to appear together, are linked by the causal
relation' - such is
116 THE CAUSAL THEORY OF TIME

the defmition of Reichenbach, who further believes that cause and effect can be
distinguished by introducing a supplementary condition. Disregarding the
condition for the moment, let us try to specify precisely what it means for two
events always to appear 'together' - is not the notion of simultaneity, or at the
very least that of a certain spatio-temporal proximity, which the author promised
to avoid, implicitly contained here? I must confess that I am unable to attach a
meaning other than spatio-temporal proximity to this word. That certain authors
(who have familiarized themselves with Minkowski's scheme of a four-
dimensional universe to such a point that they seem to have for-gotten the most
familar features of our sensible universe) affirm that the invariant spatio-
temporal relations are somehow prior to their temporal and spatial components
with regard to the choice of a system of coordinates, does not seem to me
adequate for the real cognitive process, which starts with the components and
ends up with the invariant relations.
"Thus, in abandoning the intuitive knowledge of time you will never arrive at
a satisfactory defmition of the causal relation, which was completely clari-fied
by Hume's classic analysis, which stripped causality of the metaphysical, logical,
and psychological appearances which formerly belonged to it, and which you are
bringing back. [For a detailed analysis of Hume's defmition of causal relations
and a proof of its ultimate untenability, cf. below, Part IV -
H. M.J Hume proved that the relation of cause and effect consists in a regular
succession, and is nothing more than this empirical relation, neither more
mysterious nor more comprehensible than other relations which are induc-tively
established by a certain number of experiences, and which can always be
questioned by further experiences. The apparent logical aspects of this relation,
which led to the belief in the possibility of an a priori deduction, rested on the
confusion of causality with the logical prinCiple of sufficient reason, and were
only apparent (but the fruitless efforts to deduce the princi-ple of causality
continue to this day.)68 The psychological approach, which brought about, in a
dogmatic anthropomorphism, the identification of the feeling of constraints
accompanying certain human acts, and transsubjective relations of phenomena,
was similarly misguided as to the metaphysical approach which tried to give the
empirical relation of regular succession a more or less metaphysical
interpretation in terms of necessity and action.
"Hume's analysis seems to me to be conftrmed by 'continuist' physics. The
prototype of all physical action, according to present-day science, is, as
Reichenbach points out, an electromagnetic disturbance traveling through space
with a finite velocity. Let us consider a simple schematic example to show that it
is possible to break down this continuous propagation into regular
THE WORK OF REICHENBACH 117

successions. We know that any local disturbance of an electromagnetic field is


propagated in spherical waves traveling with a velocity determined by the
constant which apears in the 'wave equation'. Thus the state of the center of the
disturbance at a given instant can be found (leaving aside the change of intensity)
after a time interval ~ , at the surface of a sphere with radius r and its center at the
origin of the disturbance. Thus the instantaneous state of the center is regularly
followed by a similar state on the sphere of radius r at the instant (t + ~ ),
provided the space under consideration is not subjected to other electromagnetic
influences. We see, then, that the propagation of waves is easily interpreted as a
regular succession of physical states. It is true that we often speak as if the earlier
state, the disturbance at the center, engenders the later states on the concentric
spheres. But if we adhere strictly to the sense of the wave equations, we cannot
deny that the expressions 'engender a state X' and 'be regularly followed by X' are
exact synonyms. This becomes even clearer if we consider the reversible nature
of the equation, whose form does not change if we substitute '-t' for the variable
't'. There-fore it is theoretically possible to determine a process, exactly the
inverse of the one just described, where a disturbance, starting from a sphere,
would be propagated to its center, 'engendering' there the corresponding state.
Yet, what has changed? The sign of 't', i.e., the direction of succession. We will
conclude, then, that the physical laws describing the gradual step-by-step
propagation· [by contact] of energy processes can be broken down into regular
successions, and offer a perfected kind of causal law , conforming with Hume's
analysis."

"What makes physical causality appear to be a more intimate relation than


that of regular succession is precisely the principle of step-by-step action [by
contact] , important to Leibniz, and since Maxwell and Faraday, one of the
directing principles of scientific thought. The old maxim underlying this
principle, that a body cannot act where it is not, does not appear to be translatable
into the language of succession: the assertion that the states of one body cannot
be regularly followed by the states of another body at a distance from it does not
carry with it the same air of self-evidence, and even seems false. Yet, basically,
this principle expresses only the fact that the physical state at a point depends
only on what takes place in its spatio-temporal surroundings, or, more exactly:
given the initial state of a place and all that occurs on its boundary during the
interval of time in question, we can deduce everything that occurs within the
boundaries during this interval. It is precisely Jordan's statement which permits
the application of the principle of causality to all regions of space, no matter how
small, and the determination
118 THE CAUSAL THEORY OF TIME

of all the states which succeed a given state. Although, as Mach had pointed
out, the causal relation is only a particular case of functional relation, it ex-
presses a dominant feature of becoming, and, as is seen by Jordan's
statement, presupposes not only the notion of succession, but the whole
topological and metrical structure of space-time as well. It follows that to
establish a relationship of physical causality, the notions of space and time
must be known in advance. Let us add further that spatio-temporal
measurements by means of rigid rods and clocks assume the use of enduring
substances. Thus, all the notions of succession, simultaneity, and duration,
which the causal theory claims to define in terms of the causal relation, seem
presupposed in the physical knowledge of causality. We might well say with
Bergson, that in the fmal analysis everything rests on the intuitions of
simultaneity and succession." 69
"Carnap's example is instructive in this respect. In his chapter devoted to
the topology of space-time,70 he assumes the notion of causal action
('Wirkungsbeziehung') to be known and defines the spatio-temporal order in
terms of it. In the following chapter, devoted to the definition of the causal
relation, he takes as known all the topological and metrical notions of space-
time, and certain others besides, to derive the causal relation. It is hardly
surprising then that, having inserted the whole temporal order into the notion of
causality, he should rediscover it there. It is true that Carnap calls the causal
relation 'determination' when it is defined in terms of space-time, and 'action'
when it serves to define space-time, and that in this last case he takes account
only of substantival action transmitted by successively contiguous material or
energetic particles. But for me, the causal relation is precisely Carnap's
'determination' with all its spatio-temporal implications."
"Furthermore, it suffices to glance at the physicist at work to be
persuaded that at no time can he do without intuitive judgments in which all
the funda-mental temporal relations play their roles, not just that of spatio-
temporal coincidence, as Reichenbach's causal theory would have it. It would
be giving in to a misleading appearance to believe that a physical experiment
could be reduced to the bare reading of dials, Le., of coincidences on suitably
arranged apparatus. There is no piece of equipment, no matter how
ingenious, which can take the place of the intuitive perception of the
relations of succession and simultaneity. Its construction has certainly
required complicated physical operations, and an understanding of the
principle of its activity, without which its handling would not merit the name
of physical measurement, also presupposes these intuitive factors. The
renunciation of which Reichenbach speaks is therefore unrealizable."
THE WORK OF REICHENBACH 119

Thus might the adversary of the causal theory speak, passing in silence over
the relativity 0 f time, which, as we have seen, is admirably accounted for by this
theory. We feel it is impossible to disagree with what he says about the negative
result of Hume's classic analysis 71 (including the supplementary conditions
resulting from Jordan's statement) as well as about the fundamen-tal role of
perceptions of time. Notice, however, that all of these perceptions deal with a
local spatio-temporal structure, an infinitesimal region of the uni-verse. If we
wish to extend these local data to more distant regions, beyond the hie et nunc of
the observer, we must necessarily have recourse to causal laws, which we
discover by local observations and which, being applicable to distant events,
serve to define their temporal order. This is how the laws of geometrical optics,
verifiable by laboratory experiments, can be extended with approximation to light
which comes to us from very distant regions. We know now, because of the work
of Robb and Reichenbach, that the whole spatio-temporal order can be defined in
terms of geometrical optics. [The untenability of their claim is shown in Part III
below - H. M.] What serves to extend the local order, therefore, is the application
of causal laws, them-selves .established by local observations. - We will still use
the method of messages, which permits the linking together of remote portions of
the universe. We have already seen that without the use of an irreversible process,
this method does not furnish a means of arranging phenomena in the order of
before and after. But nonetheless it does provide an invaluable device for the
establishment of reversible causal relations among distant events. If you see two
very similar things, you will not accept their resemblance as being fortuitous;
rather you will look for a common cause. Obviously there is the possibility of
error: but if you look for very strong resemblances and repeat your search a great
many times, you will certainly not err any more often than does the observer who
instinctively identifies the objects of his ex-perience. How do we identify a
murderer? By fmding a special characteristic in the suspect, a mark which we are
certain belongs to the criminal. This is not infallible, but it is nevertheless the
only method available. The probability will increase if you can follow the
criminal's action from the moment he commits the crime until he is indicted. The
same is true in the general case: the causal relation becomes increasingly probable
as you find more and more local events bearing the same mark. Thus causal
relation among events separated in space and time can be determined with
varying probability by local observations. Later we will show how this relation
allows us to establish the isotropic temporal order of events. Let us conclude,
then, that the causal theory of time does not force us to renounce entirely our
intuitive knowledge
120 THE CAUSAL THEORY OF TIME

of space and time; rather it maintains that it is the principle of causality


which allows us to extend this purely local knowledge to more remote regions of
the universe. In other words, causality is the link between the infinitely small
regions accessible to direct intuition; causality forms, so to speak, the unity in
the plurality of beCOming.
One might ask: if, by putting all the infmitely small regions of the universe
end to end, we might not be able to dispense with all causal considerations
concerning the temporal order. We think not. It would be like abandoning the
thermometer because within a certain interval the sensitivity of the skin replaces
it. Intuition is valid only for infinitely small regions and for velocities below a
certain limit. Just as high temperatures are inaccessible to sensibility, so world-
lines traversed at velocities close to that of light could not be acces-sible to
human beings. In such cases, the causal defmition becomes necessary.
But in deriving the temporal order of widely separated events from their
causal relations we run the risk of being the victims of Russell's humor: "We are
not likely to find science returning to the crude form of causality believed in by
Fijians and philosophers, of which the type is 'lightning causes thunder'. It can
never be a law that, given A at one time, there is sure to be B at another time,
because something may intervene to prevent B."72 And Russell con-cludes from
this that every causal law (with the "exception of quantum laws, whose
interpretation still presents difficulties) must have the form of a differ-ential
equation. Notice, however, that the limiting nature of the velocity of light tips the
balance in favor of the Fijians and philosophers: we can choose events A and B,
assumed separated by a fmite interval of time, in such a way as to render any
'intervention' impossible. For example, let A be the state of a material sphere at
time t, and B the state of the center of this sphere before following A at time (t
+ ; ) according to a causal law (c being the speed of light, and r the radius of the
sphere). Once A takes place, nothing could prevent the production of B, since
this would require an obstacle traveling with a speed greater than c. We readily
admit that neither the philosophers nor the Fijians ordinarily take these
precautions in establishing a relation of cause and effect; but Russell's argument
is none the less refuted, at least in principle. - Here, moreover, is another
refutation, less rigorous, but also less remote from practice. According to Jordan,
the fmal state of an arbitrary volume is always determined by its initial state,
provided we know the boundary conditions during the interval of time between
these two states. Since these conditions ordinarily vary very little during
experiments, we generally risk the hypothesis that they remain relatively
invariant, and thus infer the connection between the initial and fmal states with a
probability
THE WORK OF REfCHENBACH 121

sufficiently high for normal purposes. Obviously, they are never rigorously
invariable, and phenomena such as the passage of very high energy rays
through the walls of the laboratory are always possible. To combat this, well-
known experimental methods are used, such as varying the conditions of the
experiment.
If the above remarks on the causal generalization of intuitively perceived
time are correct, the relation between objective time and causality must be
considered in some way other than Reichenbach's. This extension,
comparable to the extension of the notions of color, temperature, force, and
number, is the application of a logical process which lies at the very basis of
science itself, and which may be called indirect generalization. Let us take an
ex-ample. Within a narrow interval, temperature has a determinate
psychological meaning. Science makes a non-thermal series (in the literal
sense of 'thermal') correspond to these sensations, and this series is found to
be applicable in many cases where the primitive notion would no longer have
any meaning. The physical concept of temperature, defined, for example, by
the expansion of a body, is an indirect generalization of temperature in the
psychological sense of the word: this is a generalization, because very high
and very low temperatures are defined only physically; but the intuitively
perceived temperatures are generalized only indirectly, since they themselves
do not form part of the enlarged physical series, but only the moderate
physical temperatures which correspond to them. It is again the same process
which allows the physicist to speak of the set of electromagnetic colors, of
which the visible colors form only a minute part; the same is true, mutatis
mutandis, of the many important mathematical 'generalizations', e.g., the
successive 'generalizations' of the positive integer in algebra. The legitimacy
of the process hardly seems open to question; what is at issue here is its exact
interpretation. Is it not an abuse of language to generalize as in the examples
of heat, light, time, in such a way as to make these words lose their primitive
sense? This is only a question of terminological preference which one can
pose in any application of indirect generalization, although one would be
inclined in some cases to affIrm the identity of the notions, and in others to
deny it, since the difference between the primitive series and the segment of
the generalized series which is supposed to correspond to it varies from case
to case. Thus, lOgicians thought for a long time, for example, that the number
'one' was simultaneously a member of the sets of whole, rational, real,

+" and complex numbers, in other words, that the expressions '1', '+1', '

'limn ..... 00 vn,' '1 + 0 • i' all designate the same mathematical entity. But
the 'theory of logical types', originated by Russell, has shown that this is not
122 THE CAUSAL THEORY OF TIME

a matter of identity, but of a one-to-one correspondence between the set of whole


numbers and certain parts of the sets of generalized numbers. This distinction is
of no importance to the mathematician, since all the rules of calculation remain
the same in either case. But in the eyes of the epistemo-logist, who analyzes the
meaning of the symbols, or of the metaphysician, who would be concerned with
the reality of number, the distinction would seem decisive. The same is true of the
concept of time as generalized by the causal method: that one state of a star has
taken place before another, or that one lamp in his laboratory was lighted before
another, will always involve the same meaning of 'before' and 'after' in the eyes of
the physicist. The causal hypothese~ about the propagation of light which
necessarily intervene in the fIrst case but not in the second, bring out the
epistemological difference between the two. Notice that the application of the
causal method to the second case is avoidable, but not impossible, since the
generalized notion, as such, must be applicable to every case. The importance of
the primitive notion varies from case to case: it can be rendered superfluous, at
least theoretically, by the generalized notion, or it can conserve a practical value,
or, fInally, it can remain inevitable. Thus the sense of hearing has been rendered
super-fluous for the study of the vibratory motion of air; the sense of temperature,
although theoretically superfluous, (it is possible, as has been mentioned, to
construct the whole of thermodynamics by studying the expansion of bodies) is
not so in experimental practice - it is impossible to measure the temperature of
every body that one must touch with a thermometer, just as it is impossible to eat
only what a chemical analysis has declared beforehand to be inoffensive. Finally,
in its generalization, the positive integer at no stage leaves our thought, as it is
always necessary for the counting of sets. It is this last case which applies to the
relation between perceived time and the causal order: if our analysis is well
founded the fundamental intuitions of simul-taneity, succession, and duration,
although of purely local importance, cannot be rendered superfluous by the causal
generalization.

Let us note again that the primitive and generalized series always remain
distinct independently of their cognitive role. This is not always easy to ascertain:
in the case of temperature, the confusion of our sensation of heat with the
temperature defined by the expansion of mercury, or by any other experimental
or theoretical procedure (e.g., absolute temperature) has been rare. But in the
case of numbers, the distinction between the positive integers and the generalized
numbers went unperceived for a long time, and became apparent only after a
thorough analysis. The same must be true of colors, since some eminent thinkers
still maintain today some ill-clefined identity
THE WORK OF REICHENBACH 123

between the perceived color and the corresponding vibration (Bergson). The
same is also true of objective time, which results from the superposition of
the causal order of events upon the fragmentary order of time in the literal
sense of the word; and the distinction between these two series is precisely
the fundamental theme of the causal theory of time.

4. CRITICISM, CONTINUED AND CONCLUDED: CONVENTION AND


REALITY IN THE TEMPORAL ORDER

We have seen how Reichenbach tried to draw a precise boundary between the
real and conventional elements of the temporal order - the real elements being
invariant succession and simultaneity, i.e., the topological notions, definable by
the method of messages, and the conventional elements being relative succession
and simultaneity, which imply measurements. This conventional
character finds its precise expression +
in Einstein's
wherefactor€isarbitrarilytakentobe,whilethecausaltheoryoftime the
first definition,

prescribes only that 0 < € < 1. Reichenbach seems to be in perfect agreement


with Einstein in declaring that the scientist is free to define simultaneity
according to his needs and, in particular, in such a way as to guarantee ex
definitione the constancy of the speed of light. It is precisely this point in
Einstein's work which has seemed decisive, from the epistemological point
of view, to several interpreters of relativity, according to whom Einstein's
definition would have completely changed the role of the physical concept,
which, from now on, must be identified with the set of physical operations
that serve to define it. 73
It is surprising that no one has pointed out that Einstein's purported definition
is not a defmition at all. It amounts to saying that two events not spatially
coincident are simultaneous if two light rays, originating in each event and
traveling toward each other, meet midway between them. This meeting of the
two rays is undoubtedly a sufficient condition for the simul-taneity of the two
events. But is it a necessary condition? Cannot two events be simultaneous
without there being two light rays which leave them in this or that direction? Is it
impossible for the state of a body enclosed in a perfectly opaque container to be
simultaneous with some event taking place outside the container? No one would
maintain this. In that case, what does it mean for two events not joined by light
messages to be simultaneous? Einstein's definition does not explain this. Now, a
definition must determine both the necessary and sufficient conditions of the
defined term. It follows that Einstein's famous explanation is not a definition, but
only a criterion of
124 THE CAUSAL THEORY OF TIME

simultaneity. A defmition is in a sense arbitrary, but a criterion is not. It is


therefore not possible to bring up the Einsteinian defmition in support of a
conventionalist conception of simultaneity, as Reichenbach does.
One might reply that it is not at all necessary for two light messages to
actually leave; it is sufficient that it be possible to send them so that their
departures coincide respectively with the two events, and their meeting with
a point equidistant from these events. But we have already said in connection
with Robb that the only way to make this vague notion of possibility precise
is to consider it synonymous with conformity to the laws of physics, and that
this would imply a vicious circle in the definition of simultaneity, since this
concept figures explicitly in the formulation of most physical laws. To avoid
the vicious circle, we would have to take one of these laws as defming simul-
taneity (in the method of causal decompositions, the principle of determinism
plays this role), but this would amount precisely to replacing the Einsteinian
defmition with another. - To preserve this defmition, one could of course
admit the fiction that at each instant a light ray leaves each event and (after
some suitable detours) arrives at every event which follows the first in an
invariant way. But it does not seem admissible to us to base the fundamental
concepts of science on fictions: these would suffice, of course, for the de-
rivation of any desired consequences, such as the Lorentz formulas. But
undesirable consequences would also follow. Indeed, logic teaches that a
false proposition leads to arbitrary consequences.
Einstein's first defmition appears more general. It defines the synchronism of
variously located clocks by the method of messages and deduces the simul-
taneity of events from a direct comparison with neighboring clocks. This method
allows us to dispense with the fiction of messages traveling between any two
non-simultaneous events. But it introduces another: the synchronism of clocks
cannot explain the simultaneity of two events which are not near any clocks.
Naturally, we could imagine that the entire universe is filled with clocks, but in
our opinion this is just another inadmissible fiction.
Certain authors seems to interpret the relativity of simultaneity in a more
general way: they use neither light signals, nor clocks, but emphasize the fact
that every simultaneity is measured by an observer who is stationary in a
certain system of reference, and that the measurement is therefore valid a
priori only for this system. But precisely what does it mean to say that two
events are simultaneous 'in' a system of reference? Are they events whose
simultaneity has been observed by an observer attached to the system? But then
the simultaneity of unobserved events would remain unexplainable. Shall we say
that simultaneity, just like every indication of a spatial position,
THE WORK OF REICHENBACH 125

implicitly presupposes the choice of a system of reference? We believe not. The


relativistic laws of physics would remain the same, even if all the matter in the
universe were to be transformed into radiant energy - a possibility which science
in its present state does not seem to rule out. In this universe there would be no
frames of reference, since these are material, but the temporal order of succession
and simultaneity would always remain, and would continue to be relative. But
relative to what? Our reply is, clearly, to a causal decomposition of becoming.
The relativity of simultaneity is equivalent to the multiplicity of causal
decompositions. Relativity with respect to a frame of reference is a special case
of this relativity. In the physical sense of the word, just what is a frame of
reference? It is a set of three approximately rigid rods whose simultaneous
positions determine an orthogonal trihedral angle. The property of a material
system by virtue of which it forms a frame of reference depends then on the
defmition of simultaneity, i.e., on the choice of a causal decomposition. Clearly, a
set of rigid rods can constitute a frame of reference in one causal decomposition
but not in another. Two events are simultaneous with respect to a frame of
reference formed by three rigid rods if they are simultaneous with respect to the
causal decomposition in which these rods form a frame of reference. The
relativity of simultaneity therefore rests on the multiplicity of causal
decompositions, and its discovery by Einstein was basically the discovery of a
new causal structure of becoming. Its form as a conventional defmition is only
apparent, and the demarcation traced by Reichenbach between the real and
conventional elements in the temporal order does not seem to us to be correct. We
do not deny that there are conventional aspects to this order nor do we side with
either the idealistic or realist conception of time. We merely affirm that the
epistemological problem must be posed and resolved as a whole, and that the
distinction between the topological and metrical components has no bearing upon
it.

5. THE SEMI-CAUSAL, SEMI-STATISTICAL THEORY

Reichenbach himself is not satisfied with the method of messages. In his paper
'Die Kausalstruktur der Welt und der Unterschied zwischen Vergangenheit und
Zukunft' (1925) [translated as 'The causal structure of the world and the
difference between past and future' (1978)], where he attempts an interesting
synthesis of the causal and statistical points of view in the interpretation of the
laws of the universe, he uses a more general relation than that of deterministic
causality to explain the genesis' of the temporal order, a relation which he
proposes to call 'statistical implication' (Wahrscheinlichkeitsimplikation).
126 THE CAUSAL THEORY OF TIME

This relation obtains between events A and B if, supposing A to have taken
place, one can infer with some non-zero probability the occurrence of B (this
must not be considered a definition stricto sensu of the probability implica-
tion, this latter being a 'primitive term' in Reichenbach's theory, not explicitly
definable, although defined in a certain sense by the ten axioms which it is
supposed to verify). Thus, for example, from the fact that the barometer rises we
can conclude with a certain probability that there will be good weather in the
neighborhood. Event A, the rise of mercury in the barometer, is linked to event
B, characterized by a certain state of the atmosphere, by the statistical
implication, which we express symbolically as 'A :3 B'. This notation is to point
up the fact that statistical implication is a generalization of logical implication in
the sense of Russell and Whitehead. This latter obtains when we infer with
certainty, i.e., with probability equal to one, the realization of
B from the realization of A. Thus, logical implication would be at the same time
a limiting case and a special case of statistical implication.
Having proposed this, Reichenbach proceeds to deduce the temporal order of
events from certain properties of generalized implication, which would prove that
the temporal properties of becoming are a structural particularity of the statistical
interrelation of events. Several possibilities present them-selves. For example, if
event X statistically implies event Y, while the converse does not hold, we will
say that X is earlier than Y. Or, if A and B form the total cause, and C and D the
total effect, we would have the following im-plications: A.B :3 CD, CD :3 A.B, C
:3 A.B, D :3 A.B, but the relations A :3 CD, B :3 CD 74 would in general not be
verified. Thus, to infer the future, it is necessary to know all the determining
conditions, the total cause, while inference of the past would require only
knowledge of some of the determining conditions, of the partial effect. The
difference between the past and the future would thus be defined in terms of
probable implication.
We will not cite Reichenbach's other formulations, in which the anistropy of
time is linked to the properties of statistical implication, and which seem to us to
be open to objections similar to those we shall raise in connection with the two
statements just cited. To fix our ideas, let us recall the Kantian example of the
apparent asymmetry between the inferences regarding cause and effect:
according to Kant, if I put a lead ball on a cushion, the depression in the cushion
will be determined by the smooth round shape of the ball; but if the cushion
already has a depression in it (received on some unknown occa-sion), I will not
be able to infer the existence of a lead ball. Lechalas correctly
points out that the trouble with this statement is that "the name of cause is given
to isolated objects and not to the set of determining conditions." 75 In
THE WORK OF REICHENBACH 127

the Kantian example the cause consists in a reciprocal movement of the ball
and cushion, while the effect is only the depression in the cushion, or, more
generally, the cause concerns the material state, cushion-and-ball, the effect
involves only the cushion. This last is obviously not the total effect of the
cause, since the entire state of the system ball-and-cushion is determined by
its earlier state and determines it in turn. We infer the total or partial effect
from a knowledge of the complete cause, but we can also infer the total or
partial cause from the total effect - even if there is a causal anisotropy of
time, since it is the univocality and not the symmetry of the causal relation
which comes into play. The inference is equally certain, whether it is from the
total cause or the total effect; it is equally defective if it claims to recon-struct
the complete cause from the partial effect, or the total effect from the partial
cause. In this matter the past is in no way privileged with respect to the
future. If historians appear to be better versed than prophets in their
respective domains, to such a point that history is considered a science while
prophecy is always suspect and often provokes raillery, it is because a predic-
tion is in general directly and rigorously verifiable, whereas historical state-
ments never are. It is also because the mind turns quite naturally to the past,
perceptions being spontaneously accompanied by a train of memories but not
anticipations. This anisotropic feature of the psychic life is undoubtedly a
curious fact still waiting to be explained. But it is doubtful that it should tell
us much about the nature of the transsubjective physical temporal order,
where we fmd nothing to indicate the existence of a factor like memory,
which would entail a radical anisotropy of physical duration totally different
from the statistical anisotropy stemming from irreversible phenomena.
Let us also add the singular conclusion to be drawn about the causal
relation from Reichenbach's defmition of it and the application of his axioms.
We have just seen that a pair of events, A, B, constitutes the total cause of
another pair, C, D, if the following conditions are fulfilled:
(1) The simultaneous realization of A and B statistically implies that of
CandD.
(2) The realization of A or B alone implies neither that of C nor that of
D.
(3) The realization of C or D alone implies that of A and B.
Given the transitivity of causal implication, we see immediately that the
realization of CorD alone statistically implies the simultaneous realization of
both C and D, regardless of what A, B, C, or D may be, i.e., the realization of
one part of any efffect allows us to infer the realization ofthe whole effect.
128 THE CAUSAL THEORY OF TIME

But, since any phenomenon can be considered as the effect of a suitably


determined cause, we can show from this that the realization of a part of
some arbitrary phenomenon statistically implies the realization of the entire
phenomenon, which is clearly false.
Here is another consequence, which is bizarre, to say the least.
Probability relations are applicable not only to physical events, but to all
kinds of entities, for example, to pure numbers. If the temporal order were
only a certain property of statistical implication, it should be found wherever
the latter applies, which would entail the existence of temporal relations
among the pure numbers.
Notice fmally the serious logical difficulties to be found with this theory.
Reichenbach did not take the care to specify the logical type of the variables
linked by statistical implication. Are they propositions, as one would think if
the implication of Russell and Whitehead is a special case of statistical
implica-tion? But then the example cited by the author (such as the
movement of the mercury in a barometer, or the state of the atmosphere at a
given instant) would be incorrect, as would all the statements about the
relative frequency of cases where these propositions are realized: a
proposition contains no 'real' variables and admits of no 'cases'. Let us add
that if they were propositions, the properties of statistical implication would
entail the existence of proposi-tional functions other than 'truth functions' (in
the sense of Russell and Whitehead) whose existence is doubtful and whose
use is always suspect. On the other hand, supposing that the variables range
over classes of events, unacceptable consequences are easily derivable from
the axioms of statistical implication.76

6. THE BRANCH-HYPOTHESIS [added 1975 - H. M.l

In his posthumously published book, The Direction of Time (Berkeley, 1956),


Reichenbach makes yet another attempt to establish the anisotropy of time.
But this fmal attempt seems also, as will be shown, to have resulted in
failure.
Let us discuss ftrst a difficulty which seems inherent in a deterministic
foundation of science and which preoccupied Rei<.:henbach so much that,
prior to the emergence of quantum theory in 1925-26, he actively took an
interest in an indeterministic physics, hoping that the philosophically unbear-
able difftculty which he could not help seeing in a deterministic physics
would disappear as soon as fundamental deterministic physical theories were
replaced by more powerful indeterministic theories (which, moreover, would
have a monopoly on making human life meaningful). Actually, the quantum
THE WORK OF REICHENBACH 129

theory created by Heisenberg and Schrodinger turned out to be incompatible


with determinism, according to a well-known theorem of von Neumann. Thus,
the philosophical trouble which Reichenbach attributed to the deterministic
nature of prequantum theoretical physics and which actually consisted in the
time-symmetrical structure of all major physical assumptions easily survived the
switch from prequantum theoretical determinism to the inherent quantum
theoretical indeterminism. The reason for the enduring trouble was, of course,
the time-symmetrical nature of quantum theory which proved that the issue of
temporal isotropy or anisotropy is independent of the issue of determinism
versus indeterminism - contrary to Reichenbach's earlier hopes.
This book, then, represents Reichenbach's attempt to deal with the prob-lem
of time's direction without regard to the determinism or indeterminism of
contemporary science.

7. CONCLUSION

Thus, Reichenbach's attempt at demonstrating a statistical anisotropy of time


seems to have failed. We have seen that his principle of marking, which implied a
causal anisotropy of time, also raised serious difficulties. Let us admire the great
effort to establish a temporal anisotropy which appears throughout his work and
which seems to proceed from a deeply rooted intui-tive belief, but let us ask of
him what he so justly asked of the opponents of the Einsteinian theory of non-
Euclidean space and time - to adapt intuition to the teaching of science, in this
case, the causal and statistical isotropy of time.

For in fact, the semi-causal, semi-statistical theory of time, which derived the
order of simultaneity from causal considerations and the order of succes-sion
from statistical considerations, seems to us to be condemned by science in its
present state. We have already said, in connection with Lechalas, that this theory
seems satisfactory in certain respects, but certainly not in all. We may overlook
its lack of homogeneity, which represents a step backward with respect to the
Kantian results. What appears doubtful today is the very physical effect on which
the statistical definition of succession is based, the existence of irreversible
physical processes, or, in other words, the validity of the principle of the increase
of entropy in a closed system. 77 There are two opposite points of view in the
discussion: on the one hand, there is classical thermodynamics, so well verified,
which maintains the existence of absolutely irreversible processes; and on the
other, there is statistical mechanics, which, starting with the work of Maxwell,
Gibbs, and Boltzmann, has gradually
130 THE CAUSAL THEORY OF TIME

invaded all of the domains of physics and which considers all of its processes as
essentially reversible. The discovery of thermodynamic fluctuations (such as
Brownian movement) has shaken the dogmatic faith invested in the second
principle of thermodynamics; but, on the other hand, the very foundations of
statistical mechanics (such as the 'quasi-ergodic' hypothesis) leave much to be
desired at present. Yet the facts are there: thermodynamic fluctuations prove that
the entropy of a closed system can very well decrease, thus undermining the
foundations of irreversibility. It seems therefore that we are forced to look to
statistical mechanics for an indication of the limits of the validity of Clausius'
principle. [The axiomatization of classical, phenomenological thermodynamics
by C. Caratheodory established the time-reversal variance of this theory too. Cf.
Part III below - H. M.]
From the point of view of the causal theory of time, it is especially im-portant
to emphasize a point brought out by Smoluchowski: statistical probability
behaves in exactly the same way with respect to past and future; it favors neither.
When Boltzmann made known his brilliant interpretation of entropy, one was
convinced that this quantity, although defined in terms of reversible mechanical
parameters, was essentially irreversible. From this point of view, shared by
Lechalas and still upheld by several contemporary authors,78 the passage of a
mechanical system from a state of lower entropy to one of higher entropy is
basically only the passage from a less probable to a more probable state; and the
principle of the increase of entropy is expressed by the statement that mechanical
systems tend in general toward more probable states, i.e., that of two states of an
arbitrary system, that which is more probable than the other will follow it in time.
The statistical theory of succession is thus seen to consist in time as a definition.
Mathe-matically speaking, it uses the fact that, since probability is an increasing
and therefore one-to-one and reciprocal function of time, time may in turn be
defined in terms of probability. But - and we owe the clarification of this
fundamental point to Smoluchowski - the theorem of the increase of prob-ability
is completely unfounded. A patient observer, watching a system for a sufficiently
long time, will see it pass through every possible state, no matter how
improbable, the only difference being that the duration of the improbable states
will be small and sometimes negligible, but not completely nil, with respect to the
duration of the more probable states. He will see the entropy of the system
decrease as well as increase, and these periods of decreasing entropy will on the
average be no shorter than the familiar periods of increase. He will see stones
spontaneously rising from the ground and removing a small amount of heat from
their surroundings; he will see a mixture of two gases in
THE WORK OF REICHENBACH 131

a container separate spontaneously, so that they come to occupy two distinct


parts of it, etc. It is true that at certain times the spectator will have the
impression of seeing an irreversible process; it will be because he fmds
himself in the presence of an extremely improbable state and only
remembers observa-tions made during a short period of time. This is
precisely the situation in which we fmd ourselves in practical life and even,
in general, in any scientific experiment; and this is why we so easily admit
the irreversibility of the increase in entropy and of so many other processes.
But in doing so we resemble, in the eyes of the patient observer, ''the first
plants of Spring, which think perhaps that the climate of the universe is
always getting warmer, since they never see the opposite change in Autunm".
79
Thus, we must conclude that asymmetrical succession does not admit of a
statistical defmition. We have seen in connection with Lechalas that it does
not admit of a causal defmition either. Must we impute this impossibility to
the causal theory of time and abandon it, as Russell does? We think not: if
asymmetrical succession admits of neither a causal nor a statistical definition,
it is because it has no intrinsic physical meaning and could disappear from the
formulation of physical laws without anything essential being changed.
Notice that the very form of variational equations, which reversible physical
laws take on, suggests the construction of a temporal order of events without
any possible distinction between past and future. Thus, Hamilton's principle
asserts that the path taken by an arbitrary material system in going from a
given initial state to a given final state is completely determined by these two
states, supposing that the integral of a certain function of the system ('the
kinetic potential') must attain an extreme value during the interval of time
separating the two states. In this formulation, it is not the future which is
determined by the past (the two ends of the interval playing perfectly
symmetrical and interchangeable roles), but rather the interval which is
determined by its boundaries. Thus, the temporal relation 'between' would be
more fundamental than that of before and after in all phenomena governed by
variational principles, i.e., in all reversible phenomena.
One may, moreover, assume a more general point of view, by taking the
relation of symmetrical causality as, by defmition, indifferently linking cause to
effect and effect to cause. If all physical laws were reversible, the asymmetrical
causal relation would not be invariant, but would alter its sense if the variable 't'
were to be replaced by '-t'; the relation valid for any choice of variable would
then be symmetrical. Under the hypothesis of irreversibility, asymmetrical
causality would be invariant, but we could derive from it a symmetrical causal
relation by the above definition. The invariance of the
l32 THE CAUSAL THEORY OF TIME

symmetrical causal relation is therefore independent of the problem of rever-


sibility. The same will be true of the isotropic temporal order, if it is defmed
in terms of symmetrical causality. To arrive at it, it would be necessary to
consider as a defmition of the temporal order a principle stating an isotropic
property of causality, e.g., the principle of action by contact. Let us consider
three events, A, B, C, which invariantly succeed one another in time. To
express the fact that it is B which takes place during the interval of time
separating A and C, it is sufficient to say that, according to the principle of
action by contact, A cannot 'act' (symmetrically) upon any event con-
temporaneous 80 with C except through the intermediation of an event
contemporaneous with B, and that the same is true of the action of C upon
some event contemporaneous with A, while B can very well act upon an
event contemporaneous with A (or C) without the intermediation of an event
contemporaneous with C (or A). We see that the principle of action by
contact leads to a relation of A, B, C, symmetrical in A and C and capable of
defining an invariant isotropic order. 81 In order to pass to the general relative
order, we defme: Y takes place between X and Z in a certain causal
decomposition if, in this decomposition, the events, X, Y, Z are simultaneous
with some events A, B, C, respectively, which succeed one another in an
order directly defmed by the principle of action by contact.
These summary indications, which will be completed in the second part of
this Essay, will suffice to show in what direction the definitions of Reich-
enbach and his predecessors must be changed in order to take the isotropy of
time into account and to defme a reversible and relative temporal order in
terms of the principles of determinism and action by contact. In a certain
respect our point of view approaches that of Hugo Bergmann, from whose
work we quote the following curious passage:

For clarification we can imagine before us a film of a rock falling to earth, a stream
flowing down a mountain, etc. We should not know how we should project the film. If the
flim were projected in one direction, then the rock would seem to rise and the stream to
flow up the mountain; if it is projected in the other direction, we would see the opposite. In
order to determine how to project the film, i.e., to determine the direction of the past .....
future, we must establish a law of causality as a basis. Indeed, we do not
need a special law for the rock and a special law for the water; rather, it is sufficient to
establish arbitrarily a law of causality, as for example: water flows downhill and thus the
direction of all other relationships of causality upon the flim is given empirically from the
film .... But a procedure for causality must arbitrarily be established ... We
establish an irreversible procedure and define what is cause and what is effect. We define
for the world that the total amount of entropy is increasing. With this [definition 1 the
direction of past ..... future is definitively established.82
THE WORK OF REICHENBACH 133

We quite agree with Bergmann that in the eyes of the physicist, only a
conven-tion can make it possible to decide in what direction the ftIm, once
developed, must be projected, umolled. But the task of the philosopher
wishing to explain the epistemological genesis of the temporal order seems
oversimplified in Bergmann's exposition; it is not the threading and
projecting, but rather the cutting and editing of the ftIm which is comparable
to the introduction of the temporal order into the universe. The fIlm consists
of a certain number of frames, and on each frame several images may be
juxtaposed. The juxta-position of images on the same frame corresponds to
simultaneity, the arrangement of the frames into a strip corresponds to the
isotropic temporal order, and fmally, the direction in which the fIlm is
threaded on the reel and projected corresponds to asymmetrical succession.
Only this last element of our temporal order is conventional (or rather
historical) in our view. To decide if two images are on the same frame, or
whether a given frame lies between two others, we must examine the causal
structure of the events ftImed, applying the principles of determinism and of
step-by-step action by contact. The convention about putting the ftIm on the
reel and projecting it, on which Bergmann made the entire temporal order
depend, comes only after these first two questions, whose solution is
precisely the fundamental theme of the causal theory of time. Naturally, it is
possible that some process accompanying all the events ftImed should have
taken place in one unique direction: pointing that out would then suffice to
resolve the last two ques-tions in one stroke. But universal reversibility
involves precisely the non-existence of such a process in the case where the
set of events ftImed coincides with the history of the physical universe. We
cannot 'agree' with Bergmann that the entropy of the universe is always
increasing, since there are also periods when it decreases; it would be useless
to substitute another process for the variation in entropy. This does not mean
that the convention proposed by Bergmann is false, which would be absurd,
but only that in adopting it we would be attributing to succession properties
other than those which the laws of physics confer upon it. Like Reichenbach's
principle of marking and his comments on statistical implication, the
convention proposed by Bergmann does not seem to us to offer a satisfactory
explanation of the temporal order.
VI

RUSSELL'S CAUSAL EXPLANA nON OF DURA nON

1. THE LOGICAL ANAL YSIS OF DURATION

In the theories which we have just considered, two important aspects of time,
i.e. succession and simultaneity, were defmed in terms of causality. There is
another aspect, which investigators have often considered as equivalent to
time, and which Bertrand Russell (who, however, does not favor a causal
theory of the temporal order, as we have seen) has attempted to explain caus-
ally - duration. We are using the term here in its Kantian sense, as signifying
the tendency of things to be extended through an interval of time while still
keeping a fixed identity. This is evidently a new aspect of time, not imme-
diately derivable from either succession or simultaneity: we can easily
imagine an energetic universe devoid of matter, where events succeed, or are
simul-taneous with, one another, but where there is no duration in the
Kantian sense. In saying that a monument endured from the twelfth to the
fourteenth century, I am saying more than that its construction and
destruction were simultaneous respectively with certain events whose
separation in time we know, and posterior to certain other events which serve
to date them. The essential point of my statement is that the same monument
filled the interval of time in question: this identity of the enduring object, the
identity of 'yesterday's monument' with 'today's monument', is a new datum
with respect to the relations of succession and simultaneity. It is true that this
identity in duration is a relation so familiar to daily experience and scientific
activity that philosophers in general do not seem to have noticed that there is
an empirical relation here, undefmable in purely logical terms, and above all,
very obscure. They had confused identity in duration, which relates distinct
objects, with logical identity (,numerical', as it used to be called), where there
is only one object. Logical identity was defined by Leibniz with all the
precision desirable: his principle of the identity of indiscernibles is really
only a definition of logical identity. He deduced it from the principle of
sufficient reason, not realizing that it was an analytic principle, the proof of
which, moreover, is very simple. Let there be two distinguishable objects,
A and B, i.e., such that every property of A is also a property of B, and vice
versa. Now A certainly has the property of being identical to A: B, therefore,
134
RUSSELL'S EXPLANATION OF DURATION 135

by hypothesis, will possess the same property and is identical to A. Having every
property in common is therefore a sufficient condition for logical identity; it is
also a necessary condition, since, if A is identical to B, they form but a single
object and are therefore indiscernible. Hence, we can define logical identity as
having every property in common; this is the procedure adopted by Russell and
Whitehead in their Principia Mathematica.
Now, in our opinion, a close semantic analysis 83 proves that identity in
duration is not this logical identity, assumed as a metaphysical principle by
Leibniz and correctly defined by modern logic. This seems obvious for an
object 'which has changed' while enduring, e.g., which has grown older. The
young and old Napoleon are not indiscernible, and consequently are two
distinct objects, although maintaining identity. But the same is true of objects
'which have not changed', since even if they had kept all their intrinsic
properties, the world around them would have changed, involving changes in
their relations to their surroundings. Even if the electron were immutable and
indestructible, its velocity and position would always change with respect to
a suitably chosen system of coordinates, and consequently, at two successive
instants, the electrons would not have the property of occupying the same
place or of moving with the same velocity with respect to this system. - We
could try to get around this difficulty and preserve the logical character of
identity in duration by giving a different interpretation to the propositions
expressing the change. These propositions can all be reduced to a
fundamental scheme: 'object X has property Yat instant t', or, more briefly, 'X
is Yat t'. By substituting particular values for the variables in this
propositional fllnc-tion we obtain, for example, 'Napoleon was emperor in
1804', 'Napoleon was defeated in 1812', etc. The problem is to know how this
function can be reduced to the normal form 'X is Y. Two solutions seem
possible: we can incorporate the variable 't' either into the subject, or into the
predicate. At ftrst sight, X is an individual, while Y is a class; what is the
semantic status of the variable 't'? Let us consider once more the sentence
'Napoleon was emperor in 1804'. We can disregard the past tense of the verb,
since this only expresses the temporal relation of the sentence itself (as a
physical event, which takes place at a determined instant) to the event which
it describes. If the same sentence had been spoken before or duing 1804, the
past would have had to be replaced with the future or the present, without the
objective event described by the sentence being affected. Thus, the sentence
can be brought into normal form in two ways: (1) 'Napoleon is (emperor in
1804)', or (2) '(Napoleon in 1804) is emperor'. We shall return presently to
the second interpretation, which seems to us to be the only correct one. As for
136 THE CAUSAL THEORY OF TIME

the first, which considers the whole expression 'Y at t' (emperor in 1804) as
representing a class, notice that it leaves the semantic status of the variables
'Y' and 't' indeterminate. Three hypotheses seem possible: (a) The variable 't' is a
function of the argument 'Y'; (b) The variable 'Y' is a function of the argument
't', this latter being of a logical type higher than that of the individual; (c) The
variable 'Y' is a function of the argument 't', this latter being considered as an
'individual' variable. In the first two cases, 't' is of a type higher than that of the
individual, and the number of instants, i.e., of the particular values of 't', is
determined by the number of individuals (it is
2n

no greater than 22 • , n being the number of individuals, and the number of


exponents being equal to the order of the logical type of the variable 't', supposing

the individual is of type (1 ». Now, under the substantialist hypo-thesis, the


number of physical individuals in the universe seems to be fmite and furthermore,
time is considered as being present in every system composed of a finite number
of individual particles; these particles do not engender any logical type with an
infmite number of elements, whereas time is also a variable continuum. As for
case (c), which leads to the consideration of the temporal variable as being of the
lowest logical type, and of instants (or time intervals) as real individuals, let us
point out that it takes us back to the idea of an absolute time existing outside
phenomena - a gratuitous affirmation which has been rendered unintelligible by
Relativity. [In the concluding considerations of Part IV, an anti-relational view of
time will be established. The interpretation of 'A and B at t', discarded above, will
then be recon-sidered.] We must conclude, then, that judgments which are
apparently about substances maintaining their identity through an interval of time
must be interpreted according to the fust hypothesis, which considers expressions
of the type 'Napoleon in 1804' as individual names. This means that an object like
Napoleon must be viewed as a set composed of the individuals: 'Napoleon at the
instant t, t',' etc., all linked by a relation sui generis which makes us consider all
these events as part of a single human biography, a relation whose causal analysis
is precisely the object of the present chapter. Before passing to the exposition of
Russell's theory, it should also be noted that this semantic analysis - with which
Russell might not agree, although it leads to a result which he takes for granted -
implies the complete assimilation of the notions of spatial and temporal presence.
Common sense and pre-relativistic physics establish a difference between these
concepts which seems unfounded. An object extended through an interval of time
is supposed to be present as a whole at each point in this interval, while an object
extended in space is
RUSSELL'S EXPLANATION OF DURATION 137

present as a whole only in that portion of space with which it coincides. In


our opinion, presence in time must also be viewed as a genuine extension, as
a plurality of distinct parts covering different parts of the temporal interval.
What distinguishes temporal extension from spatial extension (for this
funda-mental distinction between space. and time always persists, even in
General Relativity - Reichenbach has the merit of having insisted upon this)
is the nature of the causal relations linking the different parts of the
extension. 'Yesterday's electron' and 'today's electron' are just as much two
distinct objects, according to Russell, as are 'the electron here' and 'the
electron there'. The difference between the two cases lies in the nature of the
causal connection, which we shall now discuss following Russell.

2. RUSSELL'S CAUSAL DEFINITION

Russell fIrst points out that in the eyes of the philosopher the difference
between matter and the physical void is not one between two types of
substances, nor one between the absence and presence of a substance, but
simply a difference between the causal laws which determine the succession
of events in a void and in a place where a substance is said to be found. We
can even speak of a simple quantitative difference: "The invariant mass of
light is 0 and its relative mass is fmite. Wherever energy is associated with
matter, there is a fmite invariant mass m. But where energy is in 'empty
space', m is zero. This might be regarded as the defmition of the difference
between matter and empty space. 84 We should say that this explanation is
another of the type furnished by the systems of Robb and Carnap. It did not
take into account the cognitive process, where a difference in nature between
knowledge of matter and knowledge of the void might appear, but considers
science to be purely a set of propositions. It seems then that certain terms,
appearing in these propositions, could be replaced by a logical combination
of certain others, by virtue of the properties of these terms resulting from this
set of propositions. Thus, matter in this case would be defmable in terms of
mass and energy. This explanation would become illusory if it were found, as
it certainly seems it would be at fIrst sight, that in fact any determination of
mass or of energy presupposes the identifIcation of some pieces of matter; in
any case, the possibility of a definition of this kind in no way insures us
against a vicious circle in the epistemological sphere. - Notice too that
several authors (Born, Lewis, and even Russell, following him) have insisted
that no observation can be made in vacuo (all our measuring instruments are,
so to speak, material), and since the electromagnetic fIeld is consequently
138 THE CAUSAL THEORY OF TIME

inaccessible, in principle, to direct observation, it must be considered as an


auxiliary construction, simplifying the formulation of physical laws, but not
enjoying the same status of reality as directly perceivable matter. Without
entering into a discussion of this thesis, which goes against the general ten-
dency of science to consider the field as a reality in a certain sense
equivalent to matter, we shall only remark that it emphasizes an important
difference between the knowledge of matter and that of the void, a difference
which nowhere appears in Russell's explanation.
Now if the difference between a material particle and the field which
surrounds it can be explained in terms of a difference between the causal
laws which determine the succession of phenomena within 'matter' and in
the 'void', without having recourse to the notion of substance, the same will
be true, a fortiori, of the interconnection of the successive states of this
particle, which must be dermed independently of the notion of substantival
identity through time; this is the concept which will be defmed in terms of
causal notions. But let Russell speak for himself:
What is pecu1iaJ: about a string of events which physics takes as belonging to one electron, is a
character which is present approximately in the common-sense 'thing', a character which I
should derIDe as the existence of a first-order differential law connecting succes-sive events
along a linear route ... Let us take first the common-sense 'thing'. If I watch a moving object, I
have a series of percepts which change gradually, both as regards position, and as regards
qualities - colours, shape, etc. The gradualness of the change is the criterion by which I am led
to regard the percepts as all belonging to one 'thing'. But in a common-sense basis there are
exceptions, such as explosions. Science deals with these as rapid, but not instantaneous, changes
and so removes the exceptions. We thus arrive at the conclusion that, given an event x at a time
t, there will be closely analogous events at neighbouring times. We may symbolize this by
saying that, if there is an event
x at a time T, there will be, at any neighbouring time t + dt, an event: x + fl (x) dt + h(x) dt2 ,
where fl (x) is a continuous function of the time, while h (x) is determined by the second-order
differential equations of physics. The string of events so connected is called one piece
of matter. In the case of sudden changes contemplated by quantum theory, there is still
continuity in everything except spatial position and the spatial position undergoes a change
which is one of a small number of possible changes. Thus in this case also the new occurrences
can be causally connected with the old, though the laws of the connection are somewhat
different from what they are in the usual case ...
Thus the string of events constituting one material unit is distinguished from others by the
existence of an intrinsic causal law, though this law is only differential. A light wave, in this
respect, is analogous to a material unit; it differs in the fact that it spreads spherically, instead of
85
travelling along a linear route.

3. CRITICISM AND COMMENT

This lucid passage, which needs no commentary, nevertheless raises several


RUSSELL'S EXPLANATION OF DURATION 139

questions, some about the epistemological significance of the proposed


defmitions, others about technical details. First the question of principle: why
replace the notion of material substance, so familiar to common sense,
indispensable, perhaps, to the experimenter, with a group of events, more or less
artificially delimited? Russell answers that it is because the event is the only
immediate datum of experience and the only legitimate inference from these
data: what we perceive always occupies a small region of space-time and never
extends indefinitely in the temporal dimension, as do material substances; what
we infer concerning the unperceived entities of the physical universe are also
events (if, we might add, we adopt the postulate of the continuity of the
unperceived with the perceived, to complete Russell's thought). Yet, the last point
seems debatable: the inferences, or rather the physical constructions, although
only reluctantly isolated from the data of experience, do not seem to take into
account the postulate of continuity, which is essentially only a compromise
between the phenomenalism of the philosopher and the realism of the scientist. If
it turned out that the sub-stantialist construction (although as such it is foreign to
every immediate experience) simplified the formulation oflaws and facilitated
our orientation in the physical universe, the scientist would certainly not hesitate
to use it, leaving the philosopher the task of distinguishing its real from its
conventional elements. This could not deny the scientifically very important fact
of the simplificatory power of the construction. If the renunciation of the substan-
tialist construction nevertheless seems necessary to us in view of the present state
of the problem, it is because this construction runs counter to a postulate which is
indispensable in a different way from that of the continuity of the perceived with
the unperceived (which is inferred): that of non-contradiction. We have just seen
in the semantic analysis of section 1 of this chapter that the notion of a substance
which retains its identity throughout an interval of time is inseparable from that
of absolute time: it is impossible to reject the latter and retain the former. The
interpretation of physical laws in terms of events and, in particular, the
replacement of the notion of substance with a suitably defined group of events,
is, in our opinion, a consequence of the relativity of time.

We readily grant Russell the distinction between common sense's vague and
tenacious notion of substance, and the more precise and consequently more
widely applicable one belonging to science. Yet, in both case, his ex-planation of
them in terms of qualitative continuity seems to us inadequate. It is easy to
construct series of qualities, continuous in space and time, which neither
common sense nor science would consider as substances. Move a
140 THE CAUSAL THEOR Y OF TIME

ribbon of color behind a slit: the successive moving images seen by an observer
in front of the slit will not appear to him to form a substantial series, although the
condition of gradual change is rigorously met. Move an extended particle (such
as a rotating electron) so that the different points of its surface succes-sively
coincide with a fIxed point in space: the successive states of this point will
satisfy the postulate of gradual change, without forming a substantial series.

We might succeed in resolving this diffIculty by completing Russell's only


postulate of continuous change with certain considerations concerning the
surfaces of discontinuity, which seem to enter the picture in every case of
substantival synthesis. We imagine ourselves in the presence of two distinct
things, whether they are separated by empty space or by a third thing. In the fIrst
case, the instantaneous aspect of each thing is enclosed in an area of discontinuity
separating 'matter' from the 'void'; we know in fact that this passage between
matter and the void is not continuous, since the density of matter never decreases
in a continuous manner to zero. If the two things are separated by a third, there
are two possibilities, depending upon whether or not the third thing is separated
from the fIrst two by a void. The fIrst case amounts to the one just discussed. In
the second we must consider two substances touching along a line or a surface,
but still forming two substances and not just one. Clearly, this is the case only if
the surface of contact is Simultaneously a surface of discontinuity. Take the ink-
well on my desk: the surface of contact is the base of the ink-well - the color, the
chemical com-position, the dynamic properties all change there in a
discontinuous way. Admittedly, I could put an ink-well on my table which is
made of the same wood as the table; then most of the properties just mentioned
would be on both sides of the surface of contact, preserving continuity. But not
all of them: the dynamic properties would always change discontinuously at the
surface of contact. Thus, for example, the force required to lift the whole ink-well
would always be smaller than that necessary to lift a part of it separated from the
surface of contact by a thin fIlm. Let us conclude, therefore, that the thing is a
group of events continuous in time, but enclosed in space by a surface of
discontinuity. Clearly then, our two former examples are no longer relevant, since
neither the part of the ribbon visible through the slit, nor the point in space with
which the different points of the surface of the electron successively coincide are
enclosed in a surface of discontinuity.

But does this eliminate all the diffIculties? We are thinking especially about
an electromagnetic perturbation, which is also enclosed in a surface of
discontinuity and evolves continuously through them. Does not the difference
RUSSELL'S EXPLANATION OF DURATION 141

between a thing and a perturbation consist in the fact that a perturbation is


propagated spherically, a thing, linearly? Yet we could imagine a spherical
material surface which would expand during a certain period of time while
still retaining its geometrical form, such as a soap bubble; it would perfectly
imitate the behavior of electromagnetic perturbation, except, of course, for
the enormous difference in velocity. And we are thinking also of the notion
of radiation in needle-like beams (Wadelstrahlung'), which was used in an
attempt to reconcile the facts of quantum theory with Maxwell's theory. It is
true that it is precisely this radiation which led to the corpuscular theory of
photons, demonstrating that in the mind of the physicist, linear propagation
and corpuscularity are closely linked. On the other hand, the example of the
soap bubble is no longer relevant if we attribute a discontinuous structure to
matter, as seems inevitable according to the present state of science. But let
us not forget that the same physicist, who today speaks of the proton or the
neutron and seems to employ the notion of substantival duration in a sense
which is clear and defmable in terms of the causal relation, is the one who
just invented hypotheses about substantival ether, or who is about to give a
lecture on hydrodynamics, where our condition relative to a closed surface of
discontinuity is inapplicable, and the objection to Russell's defmition,
completely justified. Granted, the hypothesis of the continuity of matter
employed in hydrodynamics must one day disappear; but the physicist who
uses it today seems nonetheless to attach a certain meaning to the notion of
identity in duration implied by this theory. What does he have in mind when
he deduces from the provisionally admitted hypothesis of the continuity of
matter that each drop of water in the ocean keeps its identity through the
course of time? In any case, his concept of duration differs from that of both
discontinuistic science and common sense. To explain it, we would
undoubtedly have to consider the very principles of hydrodynamics as
implicitly defining identity in duration.

4. EPISTEMOLOGICAL REMARKS

Don't these difficulties arise because there is an irrational element in our belief in
the identity of objects throughout duration? Its origin is undoubtedly found in
psychological life, as is the origin of so many other illusions about physical
duration. With the exception of pathological cases involving per-sonality
disorders, we all have the unshakable conviction that we have always been
ourselves, in spite of having passed through the most diverse states: we believe
that it is always the same 'I', 'numerically' identical, who is sometimes
142 THE CAUSAL THEORY OF TIME

gay or sad, attentive or distracted, ill or well, old or young. We are not con-
cerned here with analyzing the epistemological foundations of this belief, whose
existence is beyond doubt: that, of course, is what is reflected in the contradictory
identity of bodies which are extended in time. It is the intimate solidarity of the
states of the same'!', separated from every other'!, by an impassable gulf, which
appears in the double postulate of continuity in time and discontinuity in space
which we have just enunciated. The physical universe tends more and more to do
away with every sensible character; sounds, smells, colors, each disappear in
turn, yielding their places to math-ematically definable entities - frequencies,
amplitudes, phases. Is not the identity of the object throughout an interval of time
a singularly tenacious vestige of this subjectivity in full retreat in a
depersonalized universe?
If this is the case, let us be less demanding of Russell's attempt. It cannot aim
at a precise explanation of either the inexplicable belief of common-sense stricto
sensu, or its enfeebled reflection in the universe of science. What it tries to do is
to ignore this belief in the interpretation of the data of experi-ence, by
substituting for it a set of rules having a precise phenomenological sense. To
formulate these rules, we will have to ask two questions of the representatives of
'naive realism' or of science: (1) how do you recognize that you are in the
presence of a thing (of an electron)? (2) supposing that you found yourself twice
in succession in the presence of certain things ( electrons), how would you go
about deciding whether in these cases it was a matter of just one thing which, as
you say, had kept its identity in the interval of time separating your two
observations, or of several different things? The represen-tative of common sense
would undoubtedly reply: 'I have no fixed rule, but I employ several, none of
which is infallible.' It obviously follows that our judgments about the presence or
identity of things are only more or less probable, without ever being certain; but
their degree of probability is suffi-ciently high for practical purposes. It also
follows that a true definition of the 'thing' is impossible, since definitions do not
include factors of uncertainty and probability, present in every empirical
judgment concerning a thing. All that can be said, in Kantian terminology, is that
several rules of objectifying synthesis are in use, and that their agreement
generally suffices in any particu-lar case to answer your two questions. Thus, for
example, if I see several colored spots, separated from one another by empty
space, I will not believe that I am in the presence of only one thing: the criterion
of continuity is therefore a useful rule, especially for objects at rest, although it is
not enough to answer the question in every class. If I see only one spot, I will
verify, for example, that there is no abnormal pressure being exerted on my eye,
RUSSELL'S EXPLANATION OF DURATION 143

something which would be sufficient to produce this impression. Here you


would undoubtedly interrupt: can I verify that no pressure is being exerted on
my eye? Does this not pre-suppose the identity of this eye, and conse-quently
the substantialist synthesis? And if, instead of verifying the conditions of
vision, I decide to verify the substantival character of the spot in some other
way, for example, by checking to see if it exerts some pressure on the body
which supports it, or if it offers some resistance to my hand which tries to
displace it, the substantival character ofthe body or my hand will again be
pre-supposed in the 'Verification. Must we conclude with Kant that the pure
concept of substance, as the condition of all objectifying experience, is prior
to it? No, if we follow Kant's statements literally on acting and permanent
substance; yes, if it is a question of the general notion of substantial synthesis
and of the affirmation that experience cannot be interpreted as some sort of
faithful and passive reading of the data of sensibility and memory. But
several distinctions which do not appear in Kant's terminology must be
made:
In this analysis of the notion of substantial duration, we must first of all
distinguish the psychologically accurate description of knowledge from the
determination of its legitimacy. De facto, the identification is instinctively
and automatically made. We identify the sensations connected with our eye or
our hand without knowing why or how. De jure, in a perfectly rational
science, this identification should be subject to fixed rules. This state has not
yet been reached by experimental physics, which borrows its instinctive
knowledge from naive realism, without preliminary analysis. Nevertheless,
starting today, we can ask ourselves this question de jure: in a rationalized
experiment (whose judgments would be expressed in terms of immediate
data), what role would be played by these substantial syntheses, which appear
to be inevitable in every experience, although they do not always concern the
same objects? I believe that only one answer would be possible: these
syntheses are pure conventions, in the sense of Poincare's epistemology. It is
by convention that I call the different groups of phenomena (which my
instinct identifies) aspects of my hand or eye. Starting from these conven-
tional units and applying the rules of synthesis, including Russell's, I come to
group events into enduring objects. This is no longer conventional, because
the events, in effect, must possess the properties (that of gradual change, for
example) assumed by these rules. What is not conventional either is the
agreement between groupings of events into objects, obtained by starting
from different initial syntheses. The situation is essentially the same as in the
problem of measurement. Here again the choice of the unit or of the method
of comparison is conventional, but the agreement among measurements is
144 THE CAUSAL THEORY OF TIME

not. Thus, certain initial substantial syntheses are (from the point of view
of epistemology and not psychology) conventions (which is closer to the
Kantian solution than to that of empiricism); but these syntheses only serve
to apply the rilles of identification to new syntheses. In our opinion, Russell
correctly asserts that these rules are of a causal nature, but without furnishing
a complete and exact account of them.
We must distinguish then, as we have already done in connection with
succession and simultaneity, between the points of view of a science in pro-
gress and of a finished science. We have just been talking about science in
progress, and it is only by placing oneself in a fmished science that we can
accept the notion of invariant mass as established and defme, with Russell,
the presence and absence of a substance at a given place by the numerical
value of this mass. We can then say that, for two events to be regarded as
'happening to the same substance', they must be joined by a continuous
world-line such that the state of each point is determined to a first approxi-
mation by the state of the preceding point. What is objectionable in this
procedure is a certain lack of definite sequer..ce: to be able to define a
physical concept chosen at random in terms of another teaches us very little
about either. What would be needed is an axiomatic construction like Robb's,
where every fundamental concept woilld fmd its place. Only then could the
role of the relation of duration in the physical representation of becoming be
determined.
Notice, however, that Russell's theory has an important consequence
concerning the epistemological connection between the relations of succes-sion
and simultaneity on the one hand, and that of duration on the other. We have seen
that in Carnap's systems the notion of duration is the most fundamental, and
serves to defme succession and simultaneity. The same is true in Reichenbach's
systems, since the identity of the messenger through time is presupposed in the
method of messages. In Russell's system, the opposite is the case, duration being
derived from the concept of causal continuity, which already comprises that of
the temporal order. Thus, Carnap and Reichenbach give a causal explanation of
simultaneity and succession, leaving duration unexplained. Russell, on the other
hand, defines duration without giving a causal explanation of similltaneity or
succession. The method of the causal decompositions of becoming allows for the
defmition of simul-taneity and succession independent of duration. Therefore, by
combining this method with Russell's we can obtain a satisfactory causal
explanation of the three fundamental temporal relations of duration, succession,
and simultaneity. This is why, in conformity with Russell's theory, it seems
RUSSELL'S EXPLANATION OF DURATION 145

preferable to us to consider the notions of succession and simultaneity as prior to


that of duration.
This conclusion also allows us to complete an important point of our criticism
of the Einsteinian defmition of simultaneity, and of its successive generalizations
by Carnap and Reichenbach. We have already seen that all these defmitions are
essentially only criteria, since they provide only sufficient conditions for the
defmed relations. Now the causal theory of duration seems to suggest that they
also involve a vicious circle from the experimental point of view. In fact, in all
these defmitions, a light ray is emitted and its 'return' or 'arrival' is observed. But
clearly, the departure of a light ray from a given point A and the arrival of the ray
at another point B do not justify the asser-tion that it is the same ray which links
the two points of space. To make sure, it would be necessary to verify, for
example, that the ray which left A for B encountered no obstacle during its
trajectory which would have deflected it from its path, or that no ray had been
emitted which could arrive at B before the one in question, etc. It is obvious that
the concepts of succession and simultaneity are employed in all of these
verifications of identity in duration, thereby involving the defmitions under
consideration in a vicious circle. This merely confrrms our opinion, expressed in
connection with Reichenbach's conventionalist ideas, that Einstein's great
discovery of the relativity of the temporal order, in the fmal analysis, clothed a
new causal structure of be-coming in the form of an arbitrary definition, which
was only apparent.

5. CONCLUSION

The causal theory of time, glimpsed by Leibniz, more closely explored by Kant,
and then formulated by Lechalas, received a decisive impetus from Einsteinian
Relativity. A clear evolutionary line begins with Leibniz, where we fmd only a
causal defmition of succession, imbued with metaphysics, then joins Kant, who
gave a causal explanation of both succession and simultaneity, which was
homogeneous and in the spirit of phenomenalism. With Lechalas, who linked
himself with Kant, the causal theory concentrates on the time of physics and
attempts to adapt itself to its laws. Carnap and Robb put the causal theory into an
impeccable axiomatic form, while Reichenbach inter-preted it in the specific
terms of physical theory, in close connection with General and Special Relativity,
whose philosophical significance he believes to be defmable precisely in terms of
the causal theory which arises from them. Finally, Russell alone gave a causal
explanation of duration.
The theories of time with which we have been dealing show important
146 THE CAUSAL THEORY OF TIME

philosophical problems in a new light and also generate serious difficulties. We


have tried to deal with these theories, while suggesting modifications to be made
in the ideas criticized, or conceptions to take their place. A systematic
development of these conceptions will be found in the second part of the present
Essay, where a new causal theory of time will be expounded.
VII

ALTERNATIVE APPROACHES TO TIME'S ARROW


[Supplement to the monograph of 1935] 86

INTRODUCTORY

During the last decades a number of monographic studies and significant


papers dealing with various aspects of the time problem have been published.
It is impossible to survey and evaluate them in this volume without trans-forming
it into an encyclopedic monstrosity. Accordingly, I chose a single crucial issue
which was already commented upon in our study of Lechalas and Reichenbach,
viz. the issue of a privileged temporal direction. I am perfectly aware that one
hardly does justice to studies of time as comprehensive as those of Costa de
Beauregard, Grunbaum, and Whitrow, if only their views of time-reversal
invariance of natural laws are explored. However, in this study of time, available
space is of vital importance. And the inevitable consequence is then the
restriction of topics to those dealing with time's arrow.
Two reasons, one social, the other theoretical, can be adduced in support of
the above restriction. Grunbaum and Whitrow happen to be followers of the
causal theory of time. But Costa de Beauregard is not. It would have been
crippling to limit the ensuing discussions of time theories to those classifiable
under the heading: 'causal'. However, time's arrow does provide a common
denominator both for the three aforementioned theories of time and quite a few
others which have been published during the last decade or so. Needless to say,
the concentration of Part II of this study on the problem of time's arrow does not
reflect in any way whatsoever on other insights into the nature of time which
these more recent studies make available to their readers.
The second theoretical reason for gearing the presentations and evaluations of
Part II to the explorations of time reversal symmetry exclusively is related to the
present trend of physical research. Since I am summarizing an investi-gation into
the scientific role of temporal concepts or into the philosophy of those sciences to
which time is crucial I have to choose contemporary rather than past science.
Otherwise I would be dealing with the history rather than the philosophy of
science, contrary to the subtitle of this work. The scene for a study of time's
arrow was set by Schwinger's discovery of the TCP Theorem, and the subsequent
discovery of parity violations. Schwinger's theorem establishes a tight functional
interdependence of time, space and matter. The
147
148 THE CAUSAL THEORY OF TIME

parity violations disclosed an unexpected 'chirality', of cosmic space. How-


ever, the experiments showing the effects of neutral K meson decay upon the
consecutive transformations of spatial directions and of generalized charges,
designed in Princeton and performed in Brookhaven, have centered the
attention of the scientific community upon time reversal invariance. The
number of experiments designed to check on this particular symmetry and of
theories outlined to account for the results of these experiments soon
achieved fantastic proportions.
The motivation behind the emphasis on time reversal invariance is
spelled out elsewhere in this book and has nothing in common with the time
sym-metry fad which is now sweeping the nation. However, this
phenomenon in the contemporary phase of scientific history is by no means
negligible. It provides the assurance that a study of time emphasizing the
problem of time symmetry will be a study in the philosophy of
contemporary science rather than a historical reconstruction of how
Lagrange, Clausius or Boltzmann may have felt about time's arrow.

1. K. GODEL AND S. CHANDRASEKHAR

In 1949, Kurt Gbdel, the world's foremost mathematical logician, published


a short, non-technical note about time in General Relativity. ['A remark about
the relationship between relativity theory and idealistic philosophy' in Albert
Einstein, Philosopher-Scientist. Ed. by Paul Schilpp. Evanston, 1949, pp.
557-562.] The note was followed by a technical presentation in the Reviews of
Modern Physics ['An Example of a New Type of Cosmological Solutions of
Einstein's Field Equations of Gravitation', Reviews ofModern Physics 21
(1949), 447-450]. I am taking this opportunity to summarize and comment on the
role of time in General Relativity. In most cases, and for understand-able
reasons, I am confming my discussion of time to Special Relativity.
Gbdel's point of departure was Einstein's field equations of General
Relativity connecting the spatio-temporal distribution of the metric and the
energy-momentum tensors. The form of equations involving a cosmological
constant was chosen. Moreover, to obtain a reasonably accurate solution, the
well-established axial symmetry of the distribution of all galaxies was made
use of. The solution obtained was static and,accordingly,no red shift
followed from it automatically. However, this feature, sometimes held
against Gbdel, could be remedied. To simplify the ensuing presentation and
evaluation, the actual form of Godel's solution will be taken for granted.
It is well known that, logically, the Special Relativity Principle holds in
c.
ALTERNATIVE APPROACHES TO TIME'S ARROW 149

every solution of Einstein's equations of General Relativity. This means that, for
every space-time point P of the physical universe and for every degree D of
accuracy, there exists a topological neighborhood of which Special Re-lativity
holds to the degree D. The idea of cosmic time in Special Relativity basically
comes to the possibility of mapping Minkowski's space-time onto a set of time-
like world curves such that any two points on every curve have a time-like
separation. This bi-unique mapping of the world into time-like curves is
obviously independent of any choice of an inertial frame of reference. The
essential feature of every GOdelian universe is the non-existence of this
mapping. However, for any pre-assigned world point there is a neighborhood
involving such a mapping, but the neighborhood demonstrably cannot be
extended in such a way as to include the entire spatio-temporal continuum.
Another feature of G6delian universes, of particular interest to their
explorer, can be described as follows: Given any local,material frame ofrefer-
ence which might happen to be used by a human observer, it is theoretically
possible to select a space-ship originally at rest in the frame and meeting
certain minimal requirements of weight, fuel and local speed such that a
human observer who would board the ship would be taken arbitrarily closely
to a world point representing an arbitrarily chosen era in the past of this
observer, e.g., an instant of his early childhood. This seems to imply that the
observer would be theoretically able to influence any of his past experiences,
which seems absurd since the past cannot be changed by any action taken in
the present. G6del considers this consequence of his solution to Einstein's
equations as paradoxical enough to warrant both the rejection of a realist
philosophy of time and a sympathetic reference to McTaggart's paper on 'The
Unreality of Time' [Mind 17 (1908)] .
In view of the realist attitude towards time which I struck throughout this
volume, an appraisal of Godel's line of reasoning is called for. Let me fIrst point
out that the possibility of influencing the past is inherent in some important
religious outlooks, e.g., in the Catholic outlook. If anybody committed a sin of
commission or omission and his repentance is sincere enough to earn him the
divine forgiveness, then the act of forgiving would not only make him innocent
regarding this particular sin, but would actually make the sin non-existent in the
past, by undoing his past action. However, the supernatural way out of GOdel's
diffIculty would strengthen rather than refute his idealistic conclusion. The
question is whether a scientifically admissible handling of the GOdel
predicament is theoretically available.
It seems to me that such a handling is actually available in more than a single
way. The best handling consists in emphasizing temporal symmetry
150 THE CAUSAL THEORY OF TIME

which is inherent in GOdel's solution since he actually provides a determina-tion


of the square of the differential Minkowski interval which coincides with proper
time intervals for events with a time-like separation. Thus, the sign of it is
irrelevant since only the square of it is involved. However, if time symme-try is
taken into consideration, then there is no intrinsic difference between past and
future relative to any local present. The ability of the human observer to board a
space ship that would take him arbitrarily closely to an instant of his past actually
amounts to his ability to come as closely as desirable to an instant in his personal
history which is causally related to what he chooses as his present. This,
obviously, is not paradoxical at all.
On the other hand, the fact that time-symmetry is conducive to a solution to
Godel's most significant paradox can and should be used as an argument in
support of temporal symmetry. In view of what was just said about the Min-
kowski and the proper-time intervals, this supporting argument is not rigorous-ly
needed for the stand on the idea of time which any inhabitant of Godel's
universes would necessarily take. It is merely a corroborating argument.
I mentioned that other ways out of the Godel predicament are available. Let
us sketch one that is not as natural as time-symmetry and looks definitely like an
argument ad hoc. One of the conditions of a man's ability to influence his own
past is boarding a spaceship whose velocity is comparable to the speed of light in
vacuo, i.e., c (the ship's speed should be in excess of c divided by the square
root of 2). Biological experience has never provided examples of living
organisms traveling with such a speed. No biological observation would
therefore be violated if living organisms were claimed to be unable to travel with
such speeds relative to a locally inertial frame of reference.
I am taking this opportunity to point out the compatibility of a deter-
ministically or indeterrninistically causal and symmetric view of time with
Einstein's General Relativity. Given any particular world pointP this view can be
upheld with an adequate approximation in an adequately chosen four-
dimensional neighborhood of the point P. A family of points pi with i varying
from 1 through a natural number N would then extend the scope of a time-
symmetrical and causal view of time beyond the initial neighboring. By re-
peating this extension of the first neighborhood a sufficient number of times we
would succeed in extending the scope of the theory to an ever 'increasing'
sequence of subspaces of the general-relativistic space-time. As far as I can see,
there is nothing in Einstein's second relativity theory to prevent these consecutive
extensions of the scope of a symmetrical and causal time theory. According, the
latter would hold in the entire cosmic continuum of general relativity.
ALTERNATIVE APPROACHES TO TIME'S ARROW 151

In the course of these consecutive extensions, several essential features of


Special Relativity would obviously get lost. This implies the possibility of
partitioning the four-dimensional continuum into spacelike three-dimensional
continua. Moreover, the possibility of biuniquely mapping the cosmic con-
tinuum into a family of time-like curves would be lost. Neither possibility is
relevant to the symmetrically causal view of time. The three axiomatic systems
for Special Relativity outlined in the present work would obviously have to be
severely qualified if a generally relativistic approach to the time problem were
given a try. For evident reasons, this attempt is not made in the present volume. I
am confming my investigations to relativistic theories with second quantization
on the assumption that 'relativistic' means 'covariant under a Lorentz
transformation'. No quantal theory with general relativistic covariance built into
it is presently available. My treatise deals with scientific philosophy rather than
with pseudo-scientific prophecy.
Another solution to Godel's paradox concerning the possibility of in-
fluencing the past ostensibly provided by Einstein's General Relativity was
submitted by S. Chandrasekhar. This scientist does not make use of an ad hoc
assumption concerning the maximum speed of living organisms nor does he
resort to the time-symmetry which was shown to take care of the paradox.
Chandrasekhar succeeded in integrating geodesic paths in universes blessed with
Godel's metric. He found that if the sign of the Minkowski interval does not
change along a closed geodesic then the circular orbits or geodesics in universes
metricized afa Godel have a geodesic maximum length such that the square of
the Minkowski interval vanishes along the maximal trajectory. If any circular
geodesic does not have this extremal value, then an observer traveling along it
will be taken into his future rather than into his past.
The only qualification to Chandrasekhar's solution to Godel's paradox (which,
according to Chandrasekhar, Einstein seems to have 'disliked' without criticizing
it) is that the solution presupposes a non-negative, or indefmite metric of space-
time. However, an indefinite metric was successfully used in quantum theory and
one may wonder whether it should be banned from General Relativity. The
solution based on time-symmetry gives no rise to a similar qualification.

It may be pointed out that Chandrasekhar's result is logically stronger than a


related finding of L. Ozavcith. This investigator tried to establish a model for
Godel's solution to Einstein's equations of General Relativity distinct from
Godel's model and not open to the objection of closure. However, Chandrasekhar
has succeeded in showing that without abandoning COdel's model, we can well
avoid its ostensible paradoxes, particularly the paradox
152 THE CAUSAL THEORY OF TIME

concerning man's ability to influence his own past, however distant from his
present.

2. G. J. WHITROW

Since we have decided to concentrate on the temporal vicissitudes of time's


arrow, it is only fair to G. J. Whitrow to consider his article, 'Reflections on
the Natural Philosophy of Time' (in Annals of the New York Academy of
Sciences 138 (1967), 422-432) rather than his treatise, The Natural
Philosophy of Time (London, Nelson, 1961). Between 1961 and 1967, the
'Princeton Experiments' were performed in Brookhaven. They dealt with the
decay of electrically neutral K mesons and seemed to refute indirectly the
invariance of natura1laws under time reversal.
G. J. Whitrow adduces and particularly emphasizes another, cosmological
argument in support of time's arrow. He states that from the standpoint of
General Relativity, there is a privileged group of observers, viz. those who
move with a speed equal to that of the mean speed of matter. This group of
observers will use the adjectives past, present and future in a specific way
adjusted to their respective frames of reference. Whitrow is prepared to grant
an ontological status to this particular use of temporal adjective. Temporal
asymmetry would then be founded on the asymmetry of the corresponding
relation of temporal succession.
In order to assess this view of Dr. Whitrow, we have to realize, first, that
it is a proposal rather than an argument. No verifiable evidence is pointed out
in support of the time-asymmetrical view. Only its consistency with General
Relativity is claimed. However, General Relativity is a set of partial differen-
tial equations without any specification of the relevant initial boundary and
asymptotic conditions. Accordingly, an infinite variety of world-models is
compatible with Einstein's partial differential equations. Some of these
models have actually been described in detail, e.g., de Sitter's model. This
model is characterized by the closure of the temporal coordinate of the
cosmic four-dimensional continuum. It suffices to refute Whitrow's claim of
an objective time provided with a privileged direction, since the latter is
incompatible with temporal closure. A model more in keeping with actual
cosmological data was obtained by K. G6del (cf. Section 1 (l48)). In this
model, closed time-like curves occur although the causal anomalies
conjectured by G6del were shown to be non-existent by Chandrasekhar.
Nevertheless, the model is relevant to an assessment of Whitrow's time-
asymmetry for two reasons:
ALTERNATIVE APPROACHES TO TIME'S ARROW 153

(1) The local closure of time-like world curves is hardly compatible with
Whitrow's claim of cosmic time asymmetry, since the latter also includes
local time-asymmetry. Granted that Godel's model of a general relativistic
universe enjoys no monopoly, yet it is certainly consistent with Einstein's
general field equations and it has the advantage, when compared with other
cosmic models of General Relativity, of being derived from realistic assump-
tions concerning the cosmic distribution of matter. Other features of Godel's
universes are discussed in Section 1 above.
(2) It seems difficult to decide in what sense Whitrow's observers who move
always with the average speed of matter are 'privileged'. Moreover, the very
existence of such observers seems dubious, since the averaging of the speed of
matter would have to be effected over an instantaneous cross-section of the world
determined by the requirement that all the elements of the cross-section be
simultaneous with each other. In General Relativity, the meaning of non-local
simultaneity is undetermined and the very existence of spacelike three-
dimensional cross-sections of the world is dubious if the ensemble of all
admissible world-models of General Relativity is taken into consideration.
Accordingly, neither Whitrow's local nor his cosmic time-asymmetry seem
tenable. This assessment is obviously compatible with a real appreciation of his
courageous attempt to determine the status of time-
symmetry in General Relativity. It is regrettable that no other features of his
classical treatise on the natural philosophy of time can be considered in the
context of this work.

3. A. GRUNBAUM AND O. COSTA DE BEAUREGARD

In 1963, A. Griinbaum published his Philosophical Problems of Space and


Time and O. Costa de Beauregard his two books related to the problem of time
(Le second principe de fa science du temps, and La notion de temps).
Griinbaum has covered a wide area of problems related to time and space and is a
follower of an interesting version of the causal theory of time. Costa de
Beauregard, however, does not mention the causal theory of time. The only deep
issue which they both explore deals with the anistropy of time. Accordingly, I
shall confme my discussion of their work to this single issue and use to this end a
reasonably complete address of O. Costa de Beauregard delivered at the 1964
International Congress for Logic, Methodology and Philosophy of Science, the
Proceedings of which were published in 1965.
The phrase 'time's arrow' was introduced by Sir Arthur Eddington in 1920.
He stressed that, if the law of entropy-increase cannot provide time with the
154 THE CAUSAL THEORY OF TIME

arrow, then nothing else can. He probably was unaware of the papers
published respectively by P. Ehrenfest and M. von Smoluchowski prior to
1920. These scientists showed that, in a closed thermodynamical system and
a fIxed time-interval, the probability of an entropy-increase is exactly equal
to the prob-ability of an entropy-decrease. A cosmological argument in favor
of time's arrow has been derivable from the Bondi-Gold cosmology. This
cosmology, however, has been refuted in the meantime by observational
fmdings. Accord-ingly, Hoyle has published a paper about time-symmetrical
cosmology. Karl Popper's idea about a non-thermodynamical source of
temporal anisotropy has been qualifIed by its author and refuted by us in Part
III, Chapter III. A theorem concerning the bi-unique correspondence between
the set of retarded electromagnetic potentials and entropy-increasing
processes has been well established. But no sound argument is available for
the exclusion of advanced potentials.
Other attempts at saving time's embattled arrow have been made by
Reichenbach, and Schrodinger. These authors admit the temporal symmetry of
entropic changes in isolated systems but suggest that such changes may be
correlated in various closed systems. The decisive objection to this attempt at
saving time's arrow is the lack of any adequate volume of supporting evidence.
The attempt might happen to be right. But since it is gratuitous, it does not
pertain either to physical science or to the philosophy of this science.
An alternative method of deriving time's arrow from observable pheno-
mena, endorsed by H. Weyl, is related to the propagation in vacuo of light-
waves emitted by a point-like source. The propagation consists in a
consecutive sequence of concentric spheres whose radii constantly increase.
The existence of similar sequences of spherical light surfaces which shrink to
a single point is never observed. One gets therefore, the impression that the
propagation of light in vacuo is an irreversible process. This view was
refuted by Einstein in 1910. He pointed out that the appearance of
irreversibility is due to the wave-theory of light. However, if light is,
legitimately, conceived as a swarm of elementary particles, viz. photons, then
the appearance of irreversibility vanishes completely.
The principal point of Griinbaum's and Costa de Beauregard's views of
temporal anisotropy is their admission that although the physical laws of
nature are invariant under time-reversal, this does not apply to particular phy-
sical facts. The class of facts includes, in particular, the initial and boundary
conditions of any physical system, and the species of initial conditions which
are referred to as 'asymptotic', for instance in Heisenberg's S-matrix theory.
The temporal anisotropy of physical facts is perfectly compatible with the
ALTERNATIVE APPROACHES TO TIME'S ARROW 155

time-reversal invariance of physical laws of nature. This can be illustrated by the


motions of the planets. The laws which govern these motions are temporally
symmetrical and coincide with the basic postulates of Newtonian mechanics. But
it is technically impossible to change the signs of the velocities of all the solar
planets, and thus realize a reversed process. This circumstance does not affect the
time-reversal invariance of mechanical laws of which Lagrange was already
aware. My restriction of time-reversal invariance to universal physical laws was
approved of by E. P. Wigner, in an informal discussion among the leading
physicists of our generation. He is not the only one who takes an active interest
in the problem of time-reversal invariance. The physical chemist, G. N. Lewis
and the physicist R. G. Sachs, also belong in this group.

In view of what has been said in the opening paragraph of this section, I shall
not discuss, at this juncture, other significant contributions of A. Grunbaum and
O. Costa de Beauregard to the philosophy of time. This concise account of their
views suffices to show the compatibility of their emphasis on the fact-like
anisotropyoftime and our claim of the time-reversal invariance of the sum-total
of established physical laws of nature. The same issue is discussed from a
quantum theoretical standpoint in another chapter of this work.

4. R. SCHLEGEL AND R. SWINBURNE

Two recent attempts at 'saving' time's arrow were made, from different stand-
points, by the above scientists. A brief discussion of both attempts should suffice
to show their failure. Let us start with an account of Swinburne's argument: To
begin with, I have to admit that it seems pretty difficult to give a consistent
presentation of his standpoint, I mean a presentation showing the internal
consistency of his standpoint. For, on the one hand, he does claim that the second
law of thermodynamics does establish the particular type of time-asymmetry
which he terms 'past/future sign asymmetry'. On the other hand, I do not know of
any other single law of nature or of a characteristic of many laws which can
account in any large measure for the existence of past/future time asymmetry.
The only law which refers to this asymmetry is Swinburne's own statement
asserting the past/future sign asymmetry where the sign asymmetry is construed
as the claim that any state of a system is normally a sign of some past state, but
not of any future state of the system.
The constructive part of Swinburne's claim of temporal asymmetry seems
synonymous with Watanabe's espousal of retrodictability in conjunction with
156 THE CAUSAL THEORY OF TIME

his rejection of predictivity. My refutation of Watanabe's position implies,


accordingly, a refutation of Swinburne's stand. As for Swinburne's claim that
the second law of thermodynamics supports his past/future asymmetry sign,
it should suffice, at this juncture, to refer the reader to the discussion of the
time symmetry of both phenomenological and statistical thermodynamics in
Part IV, Section V below.
R. Schlegel's attempt at establishing time's arrow is more conservative. He
thinks that the sum total of scientific evidence in support of time's arrow is
derivable from thermodynamics. This gives his reader a strange feeling of the
unreality of time because the stand he takes on the unidirectionality of time
implies that nothing has happened in physical science in the interval 1900-1969.
Yet, quite a few relevant developments did occur during this interval. Schlegel
presumes that the state S of a closed system s happened after the state Sf if the
entropy associated with S is larger than the entropy Sf. Smoluchowski and the
Ehrenfests have validly established the cyclical nature of entropic changes in
closed systems. A few decades later, Khinchin proved rigorously that, in a closed
system, an accurately defined entropy is as con-stant as energy. J. von Neumann
admitted that such was the case in prequantal theories but attempted to show that
the increase of entropy in closed quan-
tum mechanical systems is valid without any tacit 'chaos' assumptions. The
fallaciousness of von Neumann's arguments is shown in Part III of this work.
It was informally admitted by E. P. Wigner who stated to this writer that the
problem of time-symmetry has the same status in classical physics and in non-
relativistic quantum mechanics.
A more general reply to Swinburne's approach is related to its anthropo-
morphic nature. In stressing the humanly important distinction of past and future,
Swinburne disregards the fact that the terrestrial homo sapiens is an integral
episode of biological evolution started about two or three billion years ago by a
single or several protozoa all launched in the same temporal direction. There is
not a single natural law which would preclude the start of a biological evolution
chain on some heavenly body in the opposite temporal direction. Obviously, we
have no practical means at our disposal to get such a duplicate biological
evolutionary ball rolling. We do know, however, that the
single decisive feature in biological evolution, viz. the occurrence of genetic
mutations ala de Vries does not provide time with an arrow. More important-ly,
we are all perfectly aware of the fact that no human individual or group could
change the signs of the velocities of the planets circling the sun. This
technological shortcoming does not prevent us from realizing that planetary
motions are governed by time-symmetrical laws - a fact realized by Lagrange
ALTERNATIVE APPROACHES TO TIME'S ARROW 157

two centuries ago. For exactly the same reason, man's inability to know the
future directly, in a way similar to his knowledge of his past (based ultimately
upon the epistemological difference between perception, memory and antici-
pation) testifies to man's participation in the single process constituted by
terrestrial, biological evolution. No relevance to the unidirectionality of time is
involved in man's epistemological predicament.
The more general issue of the alleged asymmetry of biological time could be
treated along similar lines. I shall not elaborate on this aspect of the problem of
time at this juncture because, in spite of our initial decision to concentrate on
philosophical aspects of the time problem in quantum physics and psychology,
we are still confronted with the necessity of not overesti-mating the reader's
patience. Patience is a specifically temporal attitude. It must not be disregarded
in a study of time.
Schlegel's and Swinburne's stand on time-symmetry cannot be justified. But
they can easily be explained by the power of time-antisymmetrical feelings built
into our prescientific way of thinking. Once upon a time, it took a good deal of
mental effort to identify the physical stimulus of auditory sensations with
undulatory motions of matter in contrast to the stimuli of visual sensations
provided by electromagnetic vibrations. Getting rid of our instinctive feeling of
temporal asymmetry probably requires an appreciably greater and equally
indispensable effort.

5. S. WATANABE

In his numerous publications concerning the problem of time, S. Watanabe


worked out an interesting attitude towards the principal sub-issue of this problem,
viz. time's arrow. He is perfectly aware that the physical laws of nature are time-
symmetrical, provided they be considered at the microlevel and without the role
of the human agent being taken into consideration. According to him, however,
the situation changes radically when either the transition from a micro system to
an ensemble of such systems is effected, or when human active interference with
man's natural surrounding comes into the picture. These are the two factors
primarily responsible for the unidirec-tionality of time. He considers the ideas of
becoming, of changing, etc., as derivative in so far as the privileged direction of
time is concerned. For an evaluation of Watanabe's stand on the problem of time
direction, his essay entitled, 'Time and the Probabilistic View of the World', in
Voices of Time, J. T. Fraser, ed. (1966), should prove sufficient.

The two crucial issues are the transition from micro- to macro-systems and
158 THE CAUSAL THEORY OF TIME

the relevance of the human agent to the asymmetry of time. Watanabe admits
that every quantum law governing single physical systems is time-reversal
invariant. He thinks, however, that when a transition is effected from a single
quantum system to an infmite (probably Gibbsian) ensemble of such systems, the
initial, individual time-symmetry vanishes. I shall not challenge Watanabe's
claim, although it seems dubious to me. The decisive point is that temporal
symmetry is admitted by him not only for every single quantum system, but also
for every fmite assembly of such systems. Only infinite assemblies are claimed to
be incompatible with time-symmetry. But the transfinite number of quantum
theoretical systems (regardless of any specific transfmite cardinal-ity which may
be applicable) is obviously an idealization that must not be literally interpreted. I
am not referrring here to the orders of magnitude of 2 279 or 2 79 used by A.
Eddington on different occasions to estimate the number of elementary particles
which now populate the world. It seems more important to point out that ever
since the emergence of Einstein's General Relativity Theory the universe has
been construed as spatially finite. Should the number of elementary particles be
infmite, then so would be their density in some finite subspace. This is certainly
at variance with the standard assumptions of quantum field theory and high
energy physics.
A more effective argument against Watanabe's assumptions of an infinite
ensemble of elementary particles may be derived from purely methodological
premises. Let us assume, for the sake of the argument, that no observational
evidence against the transfinite world population of elementary particles is
available. But neither is there any sound observational support in sight for such
an overpopulated world. The assumption of overpopulation is accordingly neither
supported nor refuted by any relevant, observational evidence. In other words,
the assumption is gratuitous at the present phase of scientific evolution.
Obviously, human science is not infallible: it consists ofaxio-matizable sets of
propositions which are adequately supported by publicly verifiable, observational
evidence vulnerable to future, observational fmdings. A gratuitous statement,
however, is by defmition outside of the scope of science.

It may be fair to invoke a purely mathematical argument in support of


Watanabe's infmitist assumption. He obviOUsly espouses the so-called limiting
frequency approach to the calculus of probability, originally developed by R. von
Mises and underlying all major presentations of quantum physics. Even Feynman
shares this view, although he emphasizes the specific computational techniques
for probabilities in quantum theory, characterized by the inter-mediate use of
complex valued transition amplitudes. A rigorous presentation
ALTERNATIVE APPROACHES TO TIME'S ARROW 159

of the frequency theory of probability was already shown by Wald to involve


infinite ensembles. But finite approximate procedures are available in the
frequentist approach to probability as pointed out by von Neumann. In any event,
Popper's 'propensity-interpretation' and Kolmogorov's system of set-theoretical
axioms show that the frequentist approach is not inevitable.
The role of the human observer in the utilization of time-symmetrical
formulae for the computation of conditional probabilities is described by
Watanabe as both necessary and sufficient for making physical theories time-
asymmetrical. This claim is not supported by any sound argument, either, and
thus, it too is gratuitous. The relationship between physical events and
psychological decisions is explored in Part IV, below.
Another feature of Watanabe's approach to physical asymmetry of time is his
emphasis on the retrodictability of the past of a physical system on the basis of its
present condition in contradistinction to the failure of physical laws to provide for
a parallel predictivity. Neither the validity nor the import of this dichotomy, as
seen by Watanabe, seem convincing. In the history of the physical sciences,
nothing was more impressive than the accuracy of numerous astronomical
predictions, nothing more significant than Dirac's prediction of the duality of
matter and anti-matter. Most important, perhaps, was the Maxwellian prediction
of the luminous effects of periodic electro-magnetic disturbances. At bottom, the
causal laws of Newtonian mechanics and Maxwellian electromagnetism provide
for both complete prediction and retrodiction on the assumption that the present
condition of the system be sufficiently known and the boundary conditions in the
past or the future be correctly guessed. This holds also of the numerous and
significant predictions of quantum mechanics and quantum electrodynamics;
thus, the spectrum of the light emitted by a population of excited hydrogen atoms
involving a specifiable distribution of levels of excitation can be accurately
predicted. The analysis of the spectrum can, in turn, yield a retrodiction of the
previous quantum state of the hydrogen atom population. Quantum mechanical
probabilities are admittedly time reversal invariant.

Last, but not least, comes the ontological import of the prediction/retro-
diction dichotomy. Primarily, both prediction and retrodiction are epistemo-
logical rather than ontological; they tell us of what man can know about time and
a temporal universe rather than what there is in a temporal universe. This is a
severe limitation on Watanabe's time theory.
The cosmic validity of his claim (that our knowledge of temporal asym-metry
is conditional upon our ability to perform as a conscious agent in the world we
live in) is hardly reconcilable with the circumstance that the length
160 THE CAUSAL THEORY OF TIME

of man's written and unwritten history is of the order of one million years, in
contrast to the two or three billion years of terrestrial biological evolution
and to the many billion years of cosmological and astrophysical evolution.
Thus, in an overwhelming part of universal history, reliably known by homo
sapiens, there was no representative of his species, let alone a consciously
active representative. This does not prevent this colossal chunk of the past from
being temporal, nor does it provide the time in which the world evolved with
features foreign to the universal time as we know it. We would have to surrender
most of our scientific knowledge of time in order consistently to espouse
Watanabe's view of time. I, for one, fmd this price to be exorbitant.
On the other hand, if a deterministic causal theory of time is granted, the
two temporal directions pointing to the past and the future, respectively, are
completely fixed. (Cf. Part III, below). To select one of these directions, it
suffices to stipulate that any particular, single decision successfully made by
a human being temporally precedes the physical action decided upon. There
is no objection to such a defmitional stipulation, but little would be gained
thereby insofar as our knowledge ofthe nature oftime is concerned.
It may be worthwhile to note the somewhat surprising contrast between
Watanabe's concept of retrodictibility and the view held by another prominent
researcher, D. Layzer. This scientist claims that the future coincides with the
predictable whereas the past merely leaves traces in the present. His past-future
dichotomy is obviously as untenable as the one proposed by Watanabe. However,
although we are not in a position to espouse either view, we can offer the
following explanation of the contrast in terms of the professional bias of these
scientists. Layzer seems to be primarily interested in cosmology and was
probably very impressed by the unsurpassed accuracy of astronomical
predictions, contrasting with the multiplicity of the cosmological views held with
regard to the past. The quantum theorist Watanabe may feel that if we remember
the closure of a physical system, we can retrodict its past. But we cannot have
any comparable knowledge of the future closure of the system and this is why
predictivity fails. I am embarrassed to quote, at this juncture the German proverb:
Alles verstehen ist Alles verzeihen. [To understand all is to forgive all.]
PART TWO

DURATION AND CAUSALITY


VIII

THE INTUITIVE FOUNDATIONS OF THE


KNOWLEDGE OF TIME

1. PRELIMINARY REMARKS

The causal theories of physical time whose evolution we have traced in the
ftrst part of this work 8 7 try to explain the role of the concept of time in the
physical sciences. They bring out the causal character of the methods used
by the scientist to determine the simultaneity, succession, and duration of
physical events, and regard the temporal order of these events as a particular
circumstance of their causal interconnection. In this second part, we propose
to expound another conception of time which will both eliminate the serious
difftculties raised by these theories and extend the causal explanation of time
to the extra-physical forms of the temporal order.
In the social sciences and the humanities, one constantly encounters these
forms of the temporal order, whether it is a question of locating in time either
the psychological phenomena themselves, or physical phenomena related to
them. The literary critic who explains the genesis of a poem, the aesthetician
who determines the moment of the first occurrence of an idea in the artist's
mind, the psychiatrist who establishes the ftrst appearance of the patient's
delirium, even if he does not believe the disease to be psychogenic, are all
locating psychological phenomena in time. The same is true of the his-torian
ftxing the date of Joan of Arc's ftrst visions, of the criminologist who, in
weighing the responsibility of the criminal, determines the moment when
passion overcame him, or of the epistemologist of astronomy trying to
establish the origin of the 'personal equation' and explain why a certain
observer sees the passing of a star before hearing the simultaneous striking of
the clock. These examples show that the localization in time of psychological
phenomena is as important for the social sciences and humanities as is the
temporal order of physical phenomena for the physical sciences, and that an
epistemological theory of time cannot determine the role of time in scientific
knowledge without taking into account the extra-physical temporal order,
which appears in several different forms.
When, in the above examples, we suppose that the astronomer's auditory
perception occurred after his visual perception, or that a certain motif of a
poem was taken up after the poem's publication, by another poet, or that the
163
164 DURATION AND CAUSALITY

patient's delirium developed after the catastrophe which he witnessed, we use


the same word to designate the posteriority of a psychological state with res-pect
to another phenomenon, although this other phenomenon is sometimes a
psychological state belonginging to the same self, sometimes a psychological
state belonging to some other self, or a physical event. We also use the same
word to say that one physical event occurred after another. But we shall see in
the course of this study that the concepts of succession differ as they apply to a
pair of psychological states (whether they belong to the same self, or to different
selves), or to a pair of physical events, or to a mixed pair, and that these four
notions of succession, distinct and complementary, determine four different
forms of the temporal order: physical time, psychological time, inter-
psychological time, and psycho-physical time. These are the different
complementary components of universal time, which embraces all phenomena,
whether physical or psychological.
It is important to treat these components of universal time in an order which
will reflect their respective positions in the epistemological hierarchy. The
epistemological analysis of a fragment of empirical knowledge (such as the
fragment containing the knowledge of time) must consist above all in an ideal
reconstruction of this knowledge which would bring out the possibility of
successively verifying its different parts, without involving a vicious circle at any
stage: the verification of an interior phase must never presuppose the verification
of some later phase. It is evident that psychological time must be studied first,
since the laws governing the temporal order of psychological states belonging to
the same self can be verified by anyone, independently of any hypothesis on the
temporal structure of another psychological life or of physical becoming, whereas
the converse is not the case. This does not mean that each fragment of
psychological time must be intuitively knowable: there are undoubtedly large
gaps in the intuitive knowledge of the temporal order of our own psychological
states, which intuitive knowledge itself cannot fIll; and some successions or
simultaneities of psychological states belonging to our remote past may seem
every bit as hypothetical as a succession or simultaneity of physical events. But,
if there are cases where the temporal order is directly and undubitably given, it is
among our own psychological states that they must be sought, since these states
constitute the only reality which we know directly and indubitably. We shall put
off until later the task of determining which fragment of psychological time is
intuitively knowable. What is certain is that only psychological time has this
property.
The first observation we make in scanning our memory for a segment of our
psychological life in order to describe its temporal aspect is a purely
INTUITIVE FOUNDATIONS 165

negative one. Among the phenomena revealed by such introspection are sen-
sations, feelings, desires, thoughts, but not time. When I perceive intuitively that
a certain sound and a certain patch of color appeared to me at the same instant,
this instant is not given in the same way as are the two phenomena; it is not a
third phenomenon coordinated to the fIrst two. My judgment essentially
expresses that the two phenomena are simultaneous, and concerns only them.
When I say of two psychological phenomena that they are sepa-rated in time by a
more or less considerable interval, this temporal interval is not given in the same
way as the two phenomena;it is not a third phenomenon coordinated to the other
twO.88 My judgment amounts to saying that this pair of phenomena resembles, in
a precise respect, another pair of phenomena whose temporal distance serves as a
standard of comparison. Thus, this judg-ment compares a pair of psychological
states to another such pair, and its validity does not assume the existence of a
phenomenon sui generis called time. It is often said that the sensation of
psychological time is that of a con-tinous flow, a change, a renewing, and a
creation. Yet, what flows, changes, is renewed, is created, is not time, but always
a phenomenon in time.
We do not deny that there are intuitions which are essentially temporal,
and that, in particular, the cognitive acts concerning the simultaneity of two
psychological phenomena, or the equality of the intervals separating two
pairs of psychological events respectively enjoy a simple and irreducible
self-evidence. Still less do we wish to prejudge the issue of whether these
data concerning time are innate or acquired, which is a point of dispute
between psychological nativism and empiricism; the problems concerning
the psycho-logical genesis of beliefs about time form no part of this study.
All that concerns us here is the fact that time manifests itself to our inner
sense by the simultaneity, succession, and duration of psychological
phenomena, and that it is not a phenomenon coordinated to others. The
epistemological study of psychological time will consist, therefore, in
studying from this point of view the simultaneity, succession, and duration of
psychological phenomena, i.e., their temporal order.
We must not confuse the epistemological problem with the psychological problem
of time. The psychological studies of time, especially since the advent of
experimental psychology, have dealt mainly with the perception of time, or, more
exactly, with the perception of temporal relations, time itself not being perceivable. 89
It is clear that this question is only a special case of the general problem of the
psychology of time, which covers all psychological states essentially related to time
and not only the perception of temporal relations. To broaden the enquiry, one should
ask:, for example, which
166 DURATION AND CAUSALITY

psychological states are directed essentially toward the present (such as


boredom, surprise, perception), the past (remembrance, regret), or the future
(expectation, hope, decision), and try to learn what the common charac-teristics
of these types of phenomena are and what roles they play in the temporal
orientation of the human mind.90 All of these questions, which are still awaiting
systematic study, are psychological, since the essentially temporal phenomena
they deal with form only a small part of the totality of psychological states. But
we can consider psychological time from yet another point of view: instead of
dealing with essentially temporal psychological phenomena one can study the
time in which all psychological phenomena seem to be situated, and establish its
epistemology and axiomatic theory.91 What is the nature of duration, succession,
and simultaneity when these relations reunite psychological states belonging to
the same self? What are the intuitions and fundamental methods by which we
recognize the reality and structure of the temporal order of our psychological
states? How can all the properties of this order be deduced from a group of
simple properties? If these questions have been studied much less frequently than
problems relating to physical time, it is because some of them are confused in
appearance with those relating to the psychology of time, and most of them seem
to belong to the epistemology of physical time, since it seems that physical and
psycho-logical processes take place in the same, numerically identical, time;
there would then be no need to study psychological time separately. However,
this identification seems doubtful to us and, in any case, assumes a preliminary
analysis of the epistemological problems directly concerning psychological time.

2. INTUITIVE TIME AND MEMORY

In this chapter, we shall consider one of these problems, that of succession. What
is the intuitive sense of before and after? We are looking for a definition of
succession which would explain the self-evidence of our judgments about our
past experience, and above all about our immediate past. In particular, what is
the basis of the self-evidence of the proposition that we can remember only the
past? Would it be that remembrance is, by defmition, a representation of the
past? Yet it seems to us that the difference between remembrance and perception
is an intrinsic and qualitative one, and does not reside in their relation to a
foreign reality. On the contrary, it is this difference which could explain the
psychological difference between the past and the present. If I experience relief
after the disappearance of pain, it is because the feeling of
INTUITIVE FOUNDATIONS 167

relief contains a remembrance of the pain, while the pain does not certain
any memory of relief. It is this asymmetry in the structure of the two psycho-
logical states which explains the asymmetrical relation of succession
between them: psychological state X succeeds psychological state Y if it
contains a remembrance of Y.
This defmition (I), which immediately explains why we remember only
the past, raises several difficulties. Notice first that it seems to imply that
memory neither distorts nor omits any detail of our past experience. This
assumption is obviously seriously mistaken, since, on the one hand, there is a
certain distortion attached to perhaps every function of memory, while, on
the other, memory is always counteracted in part by the function of for-
getting. Forgetting undoubtedly has vital positive significance, as is shown
by the observations of psychoanalysis, and it seems probable that if a human
being were to be deprived of his ability to forget, his life would be rendered
not only more difficult, but impossible. It is therefore necessary to eliminate
this fiction of an infallible memory and fmd an interpretation for the gaps in
consciousness produced by forgetting. We will speak later of the gaps which
memory cannot fill. As for errors of memory, perhaps frequent and
inevitable, notice that the only danger they present for definition (I) is that
psychological state A can contain the remembrance of psychological state B,
although B never existed: must we say, in accordance with this defmition,
that state A succeeds the non-existent state B? I believe that defmition (I) in
no way authorizes this conclusion. It amounts to identifying the relation of
intuitive succession with the relation between the state containing a
remembrance and the state being remembered. For this relation to exist, its
terms must exist. Therefore, if the state to which the remembrance refers (as
in the case of state B above) does not exist, the defming relation does not
exist either, nor does the defined relation, i.e., succession. Defmition (I),
therefore, does not oblige us to admit the succession of non-existent states.92
How far is this defmition applicable? In other words, under what condi-
tions is a psychological state accompanied by the memory of the preceding
state? It seems that, in general, psychological phenomena do not suddenly
disappear from consciousness, since a more or less long series of memories
accompanies their production. When there has been a brief sound, one can
always catch oneself listening to the second which has just vanished. This is
not the perseverance of an impression, i.e., the perception of an increasingly
feeble sound, but the brief memory of a sound of constant intensity. It is
precisely because of these memories that I regard the sound as 'having just
been produced'. The presence of these spontaneous memories is evidenced
168 DURATION AND CAUSALITY

by the process of speech: in speaking a sentence, one makes each successive


word agree with the first word according to the rules of syntax, which shows
that the memory of the fIrst word persists throughout the sentence. The law
according to which a continuous series of memories must accompany every
psychological phenomenon was formulated by Brentano and called 'the law
of primitive association' .93 We can say, therefore, that within the limits of the
validity of Brentano's law, the succession of two psychological states can be
defmed by the asymmetry of their structures. Now, the series of spontaneous
remembrances accompanying the production of a given psycho-logical
phenomenon is broken sooner or later: therefore, definition (I) would be
applicable only to psychological states sufficiently close to one another.
Of course, Bergson has stated that each state of consciousness reflects its
entire past. But this statement can hardly be used as a defmition of succession,
since state X could very well follow another state Y without containing a memory
of it. However, it seems that in this case Y can be indirectly linked to X by a
chain of memories. It is not necessary that the feeling of relief coming from the
extraction of a tooth should contain a memory of the tooth to be considered as
following it; it is enough if it contains the memory of the pain, which in turn is
accompanied by the memory of the extraction. Thus, if in passing through state
X, I remember state Y, which in turn contains the memory of state Z, the intuitive
succession of X and Z is assured. Clearly, the number of these intermediary states
could increase. We are thus led to the following generalization of the preceding
defmition: state X succeeds state
Y if there is a fmite sequence of psychological states whose two ends coincide
with X and Y respectively and each term of which contains the memory of the
preceding term (II). It is evident that the definition (II) thus generalized permits
the arrangement, in the order of before and after, of any pair of phenomena
forming part of an uninterrupted series of psychological states. To this end it is
sufficient to link two states, respectively, simultaneous with the elements of this
pair by an intermediary chain of states such that two
consecutive states are always sufficiently close to each other for Brentano's
law to be directly applicable to them.
The same will not be true if the two states belong to two series separated by a
gap, such as my states of today and yesterday. From the point of view of
universal time, these gaps arise from events which occur between my states of
last night and today. Yet, psychologically speaking, nothing takes place in that
interva1.94 If I am able to speak of a gap all the same, it is because the normal
interconnection of memories within the series of states of one day is distorted for
the pair of states separated by the pause for sleep. I can therefore
INTUITIVE FOUNDATIONS 169

defme in purely psychological terms the existence of a gap between two


consecutive series of psychological states, with the requirement that the fIrst
state of the second series should not contain a memory of the last state of the
fIrst series. This definition does not take into account the exceptional case
where my fust thought of today is precisely a memory of my last state of
yesterday, whereby the two series intrinsically form only one, and from the
psychological point of view, there is no gap whatsoever. Apart from this
exceptional case, it will be suffIcient that at least one memory link my states.
of today to those of yesterday in order that I may consider any state of today
as later than any state of yesterday. For this purpose I shall use the following
rule: if state X is not a member of the uniterrupted series of states Sand
contains a memory of a state of this series, any state, which belongs with X to
an uninterrupted series, is regarded as later than any state of S (III).
Notice that Rule (III) allows us to bridge the gaps separating psychological
states, but not to fill them. This point becomes clear if we take into account gaps
of a kind other than those we have just been considering, which are gaps only
from the point of view of universal time. These are gaps which exist in cases
where there are many psychological phenomena in the interval between the two
series, but where these phenomena are not linked to the two series by inituitive
temporal relations. It could be, for example, that in the series of days in which we
have lived there is one which is not linked by any chain of memories to the days
which preceded or followed it. From defmitions I-III we would have to conclude
that the psychological states through which we passed during that day have no
intuitive temporal relation with the rest of our conscious life, and that, in
particular, they are not later than the states of the preceding day. This seems
absurd, if we think in terms of the general notion of succession, which allows us
to establish by indirect methods (which we will discuss in Section III) the
temporal relations among the phenomena in question. But it is no longer absurd if
we stick to the intuitive sense of psychological succession, defIned by the
interconnection of memories and alone capable of explaining the fact that we
remember only the past. We shall say that the day in question constitutes a gap of
the second kind between the days which respectively preceded and followed it.
The existence of such a gap can manifest itself in the flux of consciousness not
only by distortion in the interconnection of memories, which also appears with
the gaps of the fust kind, but also by the particular phenomena which do not exist
in those gaps. Thus, if I see an object during day A which I do not remember
having seen during day B nor before it, I can experience the characteristic feeling
of recognition, of di1ja VU, which is explainable only by the hypothesis that
170 DURATION AND CAUSALITY

between days A and B there was an instant during which I must have perceived
this object for the first time. If the two days are widely separated, the general
alteration in my entire consciousness which I experience during day A can warn
me of the gap which exists between A and B.
Rules I-III express the fact that although I necessarily remember at each
instant what I have just experienced, I can also be transported in memory to more
distant regions of my past, and that thinks to these memories - which both re-
establish the temporal continuity broken by sleep and make the connection with
my remote past - my conscious life organizes itself into days, months, and years.
Naturally, it is possible that on a given day, I shall not once recall any of my
states of the day before, but on the following day, I shall recall the states of these
two days. But here it will still be possible to arrange the states of these two days
into the order of before and after if the states of the day before yesterday seem to
be 'more distant' or 'older' than those of yesterday. What does this mean? We were
led to defmition (I) of succession by deriving the difference between past and
present from the qualitative difference between memory and perception. We
know that this difference can vary and even disappear, since there are insensible
and con-tinuous transitions between memory and perception. Of two memories
simultaneous with a given perception, one, considered as a memory, can differ
more than the other from this perception, considered as a perception. We will
then say that it is the former which concerns a more distant past (IV). Thus, in
remembering during day A what I experienced during days Band C, I will be able
to establish that my memories of C appear more different from the present
perception than do those of B, and this will allow me to locate Band C in my
intuitive time. This comparison of two memories with a simultaneous perception
can be clearly seen in those characteristic states of mind where, retracing the
course of time and immersing ourselves in our former joys and sadnesses, we
seem to relive our entire conscious past at the expense of the present, which
vanishes:

Was ich besitze, seh ich wie im Weiten, Was


ich besass, wird mir zu Wirklichkeiten.

[What I now possess, I see as if from afar,


What I once possessed, becomes my reality.]

If, while in that frame of mind, we were suddenly asked to locate in time one
event taking place in the fragment of the past which we had been reliving with
respect to some other past event (even relatively more recent), we should
INTUITIVE FOUNDATIONS 171

be unable to reply, even though both memories were there. What we would
need in order to locate these two events in time would be a change of
attitude, the return of our attention to a present perception, so that it may
serve as a standard of comparison for the two memories.

3. INTUITIVE TIME AND PERCEPTION

The considerations advanced in the preceding paragraph were meant to show


how the intuitive order of before and after, defmed for each sufficiently close
pair of psychological states in terms of the asymmetry of their structures
(i.e., only one of them contains a memory of the other), can be extended,
thanks to the interconnection of memories, to apply to psychological states
which are separated by larger intervals. But defmition (I) needs to be
completed from still another point of view: psychological state X succeeding
state Y does not contain a memory of Y, not only when their distance in time
is too great, but also when it is too small: if I perceive a sufficiently short
motion, for example, the fall of a body, no memory of an earlier phase of the
motion is contained in the perception of a later phase. Rather, the whole
motion is perceived en bloc, as a sort of temporal figure. Similarly, in
pronouncing a two-syllable word, we do not remember the first syllable at
the instant when we pronounce the second; again, it is the entire word which
is perceived en bloc. This perception implicitly contains a certain knowledge
of the temporal properties of the perceived phenomenon, since we easily
distinguish pheno-mena which differ from one another only by the temporal
order of their phases. Just how far does this knowledge go? 95
I believe that the global aspect of temporal figures is the only temporal datum
perceived. The notions of succession and simultaneity arise only when memory
enters in; they serve to describe and analyze the properties of temporal figures in
terms borrowed from the order of memories. In striking two keys of the piano in
a sufficiently short interval I produce a temporal figure composed of two sounds.
Experience teaches me that I can form two other temporal figures from these two
sounds, but when I distinguish these figures (Le., when I distinguish an ascending
sequence from a descending sequence and from a consonant pair of sounds) I do
not think of their diversity in terms of the temporal relations of their elements.
The number of temporal figures that I can form from three sounds is much larger;
but again, in intui-tively distinguishing them from one another I do not analyze
them in terms of the temporal relations linking the sounds which constitute them
- it is the global aspect of each figure which allows me to compare and classify
them.
172 DURA TION AND CAUSALITY

This classification is therefore pre-temporal; the notions of before, after, and


during come in only later and are used to analyze temporal figures which are
perceived globally.
How is this analysis carried out? When I am called upon to decide which
sound of a recently produced sequence preceded the other, I mentally re-produce
a similar sequence, prolonging, for example, one sound in such a way that, with
the perception of this sound, I come to remember the other. Therefore, it is with
the aid of memory that I introduce a temporalorienta-tion into the pair of
sensations. Obviously, this prolonging is not the only way of utilizing memory; it
is sufficient that one of the two psychological states forming the figure under
analysis be part of a third state containing a memory of the other, for the latter to
be recognized as having taken place before the former. I will know that the note
do has sounded before the note mi if the perception of the motif mi fa sol, of
which mi is a part, contains a memory of do.

These considerations show that two distinct temporal orders are associated in
the intuitive order: the time of perception and the time of memory. The time of
memory is an anisotropic order founded on the asymmetrical relation of
succession, defined in (I). The perception of temporal figures involves an
isotropic order. Thus, if a theme of three notes is played slowly enough, there
will be a time when the first two tones are perceived, and another time when the
last two are perceived, but the whole theme will never be perceived all at once.
This case, in which the two outer tones behave differently from the middle one,
permits the defmition of the intermediate position of this last, but not the
anteriority of the first with respect to the other two. It is therefore an isotropic
order which is implicitly fixed in the perception of a temporal figure. This order
is closely linked to the anisotropic order of memory: if three states succeed one
another according to the order of memory, the one which takes place neither
before nor after the other two will take place between them in perceived time.
The association of these two orders within the temporal sequence of
psychological states is comparable to the continuous order of points on a straight
line in space, which can be described using either the asymmetrical relation 'to
the left of' or the symmetrical relation 'between'. Nevertheless, in the case of
space these are merely two descriptions of the same linear order, while in the
case of time these two orders are distinct and their constant association is only an
empirical fact. If our perceptions were rigorously instantaneous and devoid of
any temporal extension, the perceived temporal order of our psychological states
would disappear and only the interconnection of memories which insure the
existence of before
INTUITIVE FOUNDATIONS 173

and after remains in the flux of consciousness. If our psychological states were
not accompanied by memories, we would live in a perpetual present free of the
shadows of the past; but these states would nevertheless have an isotropic order,
perceived at the very heart of this present. Finally, if completely instantaneous
psychological states were forgotten immediately after they occurred, time (in the
intuitive sense of the word) would cease to flow.
Perhaps it will be said that these states conserve at least the order of intui-
tive simultaneity. But is not simultaneity a limiting case of succession? It
seems easy to define a rigorous simultaneity in terms of succession, saying,
for example, that states X and Yare simultaneous if they occur 'between' the
same phenomena, or if every state which is subsequent (or prior) to one
phenomenon is also subsequent or prior to the other. However, the relation
thus defined would clearly not be intuitive, since to verify it, it would not be
sufficient to perceive the temporal figure formed by the two states: one would
also have to examine all other states of the same stream of conscious-ness
from the standpoint of succession. In order to avoid this consequence, instead
of considering all the states preceding and following X amI Y, we could
consider only those within some interval containing both X and Y; but this
would still require the examination of a large number of states, far too many
for the relation to remain intuitive. Shall we say that two psychological states
are simultaneous if neither of them is earlier or later than the other? The
existence of gaps of the second kind shows, nevertheless, that the absence of
intuitive succession between two psychological states does not warrant the
conclusion that they are simultaneous. Can we add that two states that do not
succeed each other form part of the same perceived temporal figure? This
would entail their simultaneity, if they were both strictly instantaneous. But
psychological states are never strictly instantaneous and therefore can very
well not succeed one another within a temporal figure and still not be simul-
taneous (for example, if one of them begins before and ends after the other).
Notice, however, that in this case the two states are at least partially simul-
taneous, i.e., they contain respectively two phenomena which are strictly
simultaneous. It is this partial simultaneity of psychological states which
seems to us to be intuitively given, in this sense: the analysis of a temporal
figure, carried out with the aid of memory, can show that a certain fragment
of this figure is neither earlier nor later than some other fragment and con-
sequently must be regarded as partially simultaneous with it. Like a strict
simultaneity, partial simultaneity is a symmetrical relation by defmition but
unlike the former, it is not transitive: two phenomena partially simultaneous
with a third can succeed each other in time. This circumstance confirms our
174 DURATION AND CAUSALITY

opinion that intuitive simultaneity is only partial simultaneity; we know, in


effect, that because of the approximate nature of any sensory equality (of which
partial simultaneity is an example) it is possible in each case to form a chain of
phenomena in which any two contiguous phenomena are always sensibly equal,
although the two ends of the chain are sensibly unequal. This would be
inadmissible if we were dealing with a strict equality. The non-existence of a
strict intuitive simultaneity is further confirmed by the fact that the stream of our
states of consciousness is not subject to a natural stratification; it cannot be
decomposed into cross-sections of simultaneity. When the foliage of a tree sways
in the wind, I indeed perceive the temporal figures formed by the movements of
the branches, but the simultaneity of these movements in no way seems given to
me; I feel that it would require an effort of analysis on my part to synchronize
these movements, which appear to occur alongside one another, their phases
umelated by any intuitive simultaneity.

4. THE CONTINUUM OF INSTANTS ATTACHED TO INTUITIVE TIME

We can see that the temporal order of psychological states cannot be assimi-lated
into the continuous and unidimensional order of instants, which is often taken as
constituting the very nature of time. In speaking of the stream of consciousness,
one often visualizes an infinitely divisible stream filling its bed in a continuous
way and continuously flowing toward its mouth, that is death. The extended
psychological phenomena in this stream of consciousness would then be
comparable to blocks of ice dropped into the stream at various levels and carried
along by the current. Two drops of the stream are 'simultaneous' if they are
equidistant from its mouth; one drop 'succeeds' another if it is closer to the mouth
of the stream. The relations of simultaneity and succes-sion thus defmed would
have all the known properties which are necessary and sufficient to give time the
structure of a one-dimensional continuum.
But it seems impossible to extend these defmitions to the blocks carried along
by the current, since a block does not have a determinate distance from the
mouth. Only in very special cases would the notions of succession and
simultaneity be applicable as defmed to whole blocks. Thus, if it hap-pened that
the two ends of one block and the two ends of another were equidistant from the
mouth of the stream respectively, the two blocks could be regarded as
'simultaneous'; the simultaneity thus defined, which is precisely the strict
simultaneity considered in the preceding section, would be symmetrical and
transitive. Similarly, if each point of a given block were
INTUITIVE FOUNDA nONS 175

closer to the mouth than every point of another block, the former would
'succeed' the latter, and the succession thus defmed, which is precisely the
succession analyzed in (2), would be asymmetrical and transitive. But the order
of simultaneity and succession thus defmed would not possess the fundamental
property of connectedness - there can be blocks in the same stream which are
neither simultaneous nor successive. This difficulty is magnified when we take
into account the fact that all psychological pheno-mena are extended in time: it
follows then that we must abandon the fiction of an infmitely divisible stream in
our image of the current and assume that the stream bed is completely filled with
blocks sliding over one another toward the mouth under the influence of their
own weight. This picture, although undoubtedly crude, represents more faithfully
the temporal structure of the stream of consciousness. 96 It brings out the fact that
this structure is quite different from that of a linear continuum, and that it would
be only by an appropriate conceptual construction that we could attach a
continuous one-dimensional order of instants to the intuitive order of
psychological states.
What is this construction? We ordinarily describe the fundamental proper-ties
of psychological time in terms of those of simultaneity and succession, conceived
as linking instantaneous psychological states, in perfect analogy with the order of
points on an oriented straight line. This procedure, which obviously rests on the
possibility of replacing instantaneous psychological states, which do not exist
anywhere in the stream of consciousness, with very brief states, has only an
approximate value. In order to obtain a more precise representation we could
apply Whitehead's method of extensive abstraction, which considers only
relations among phenomena extended in time, and regards the instantaneous state
as the limit of an infmite series of states whose temporal extension decreases
indefinitely.97 It seems, however, that one would thus only shift the difficulty,
since an infmite series of states is just as unobservable as an instantaneous state,
and furthermore, it is very doubtful that the temporal extension of a psychological
phenomenon could decrease indefmitely. It seems preferable to us to attach the
notion of an ideal instant to the notion of intuitive succession conceived as
linking extended states, using the well-known conception that the point can be
considered as the boundary of a segment. This statement, taken literally, does not
make much sense, since, if only the temporal segments are given, their
'boundaries' must be regarded as fictitious entities introduced for the convenience
of language and can always be eliminated in favor of constructions involving
only segments. The problem therefore boils down to that of knowing how to
translate statements which are apparently about the boundaries of a temporal
176 DURATION AND CAUSALITY

segment into statements about the segments themselves. Just to flx our ideas,
consider a typical statement of the form: 'Phenomena A and B are separated by
the instant t'. It is equivalent to the conjunction of the following two propositions:
(1) phenomenon A ends precisely at the instant when pheno-menon B begins; (2)
the instant at which A ends is the instant t. It is clear that the role of instants
would be reduced to that of segments if we were able to say what must be
understood by the proposition that a given phenomenon ends at the instant when
another begins, and that two pairs of phenomena are separated by the same
instant.
The simplest explanation of the statement, 'A ends at the instant when B
begins', would be that A precedes B, and nothing occurs between A and B, i.e.,
there is no phenomenon, C, after A and before B. However, this explana-tion
would assume that the temporal extension of phenomena can decrease
indeflnitely: if we do not assume this hypothesis, it remains possible that there is
a fmite interval between A and B but one which is too brief for any phenomenon
C to take place in it. We can avoid this diffIculty by using the notion of partial
succession. We will say that phenomenon X is partially later than Y if every
phenomenon later than X is also later than Y, but not vice versa. For
phenomenon A, which is earlier than B, to end at the instant when B begins, it is
necessary and suffIcient that no phenomenon partially later than A be earlier than
B. If, in a pair of phenomena, there is one which ends at the instant when the
other begins, we will say that this pair is separated by a single instant; in a pair
separated by a single instant, we will call the earlier phenomenon the flrst
element of the pair. Two pairs are separated by the same instant if the fIrst
element of each pair is earlier than the second element of the other. Similarly, we
will say that the instant separating pair C is earlier than the instant separating pair
C' if the fust element of C is partially earlier than the fust element of C'.

The succession among psychological instants, which was just defined by a


chain of defmitions resting on the succession of psychological phenomena, is a
different concept from this last and is endowed with quite different properties.
From the asymmetry and transitivity of intuitive succession, we can easily derive
the asymmetry and transitivity of the succession of instants; but in contrast with
the fIrst, this last relation possesses the important prop-erties of continuity and
connectedness, which determine the linear character of the order of instants: if
instant t 1 is earlier than instant t2 there is always a third instant, t3 , after tl and
before t2 . Of any two instants between tl a'nd t2 , there is always one earlier than
the other. Thus, all the instants situated between tl and t2 form a continuous
segment, and it is by ideally prolonging
INTUITIVE FOUNDATIONS 177

this segment in such a way that it embraces all the instants earlier than tl and
those later than t2 that we form the idea of a linear continuum. However, from
the point of view of the intuitive data, the idea of an inftnite time is no more than
a pure possibility. What is actually accessible to observation is only a more or
less extended series oftemporal segments, separated by pauses for sleep, with
dream phenomena floating vaguely among them.

[5. OTHER INTUITIVELY KNOWABLE ASPECTS OF TIME]

SO far, I have stressed only relational aspects of intuitively knowable time.


However, in several recent and signillcant publications an anti-relational stand
(which we shall take up in the last Part of this investigation) is implicit. We are
told for instance that "perceiving the ongoing flow of time is one of man's
immediate experiences" ,98 or that "consciousness is a presentation of an
extended manifold" 99 or that "physical time is a continuous variable, psy-
chological time is a discrete variable".lOo Another, more cautions investigator
states that "man has the capacity to resolve the stimuli into the temporal
properties of succession and duration ... These temporal properties are accepted
as realities. lol An experimentalist informs us that "an interval of 2 milliseconds
separating two sounds enables the subject to say that he hears two sounds, not
one, but he needs an interval of about 20 msec. to decide the order in which the
sounds occur." The researcher who made the last statement asserts also that
people have the capacity "to perceive time directly as a quality with measurable
and communicable units." 102
These various statements made by reputable investigators have the exciting
property of being incompatible with one another, and, in some cases of being
self-contradictory. If man actually perceives the ongoing flow of time, then
perceived time, being a flow, could not possibly be a discrete variable. These two
statements cannot both be true. But they can obviously both be false. Despite
having concentrated a good deal of time on the issue of time, both physical and
psychological, I have never caught myself perceiving the ongoing flow of time. It
seems that the 'flow of time' is a hydro dynamical metaphor which must not be
taken seriously or literally. It is true that, while concen-trating on the successive
recollections of a sound which I have just heard, I may be able to assert that the
sound gets increasingly remote from my experienced present. The content of this
experience has nothing to do with the flow of time but with the changing
memorative image of a sound I have heard. To the best of my knowledge, we
never directly perceive time, or any part thereof. But we are· directly aware of
our sensory, proprioceptive and
178 DURATION AND CAUSALITY

introspective data, whose temporal properties may be ascertained intuitively,


without any inferential process or physical clock being involved.
I am using here the word 'property' as literally construed in contemporary
logic. I am not prepared to identify properties with dyadic or four-term relations
as implied in one of the above quotations. What are the directly perceivable
properties? In one preceding quotation, a flicker was distinguished from a
succession and the physical threshold of each was indicated. I think that this is a
valuable quantitative supplement of what has been said in the preceding sections
about the perceived succession versus the perceived tem-porally extended whole.
Needless to say, the physical lengths of the two stimuli intervals are not
perceived in a similar manner.
Nevertheless, an intuitive awareness of the time interval separating two
virtually instantaneous data certainly does exist. They are either actively or
passively accessible to an intuitive appraisal. By a passive intuitive evaluation of
the length of time elapsed between two quasi-instantaneous sensory data, I mean
the well-established human ability to specify this length pretty accurately without
consulting a watch or drawing any similar, observationally based inferences. The
basis of active evaluation of a time interval elapsed between two sensory or other
experiences is man's ability to produce a similar coupling of data separated by an
appreciably equal time interval. This is one of the many intuitively accessible
aspects of time, not taken into considera-tion in the two preceding sections. The
number of these intuitively accessible aspects is considerable and may even be
infmite (if the spatial analogue can be used at this juncture.) As a rule,
topological properties of spatial, two-dimensional figures are intuitively
accessible and their cardinality is obviously transfmite. Although intuitive time is
one-dimensional and offers fewer opportunities for immediate inspection, this
limitation can be compensated for by considering simultaneous multiplicities
evolving temporally at various paces. Mozart's Divertimento, K. 517, was
written for violin, viola and violon-cello. Anyone familiar with this composition
will be able to follow each instrument separately in the course of a competent
performance. But he will also be in a position to perceive the temporally two-
dimensional patterns involving any two out of the three instruments and, more
importantly, the temporally three-dimensional pattern of the trio as a whole. The
number of directly perceivable temporal aspects of a performance of the
Divertimento is considerable. The successive reappearances of a musical motif
in the per-formance of three instruments, the quality of their simultaneous impact
on the ear, the rondo pattern of the last part of the composition, are just a few
examples of intuitively accessible aspects of the temporal manifold
INTUITIVE FOUNDATIONS 179

caused by listening to a perfonnance of this composition. It seems deplorable


that, up to the last decade, the experimental and theoretical investigators should
have concentrated almost exclusively on our perceptual knowledge of temporal
length.
One exception may be granted. Much is known and understood at present
about the time experienced by mentally deranged people, e.g., schizophrenics. It
seems probable that in the long run a study of the psychopathology of time will
benefit our understanding of temporal perceptions of nonnal adults. Cases of
psychoses which are specifically time oriented are of particular promise in this
respect. Sometimes, space and time give rise to similar devia-tions, e.g.,
chronophobia and agoraphobia. But the phenomenon of 'lapse of time' has hardly
a spatial analogue. Cases when the impairment of a cerebral region is known to
cause a disturbance of temporal perception are obviously of particular interest.

However, I have to admit that m~.:L!esearchers seem to concentrate on our


ability to estimate temporal intervals intuitively (i.e., without a clock or a similar
device being utilized). Our striking ability to wake up at a set hour without the
help of an alarm clock is often referred to as testifying to the existence of a
'human clock'. Psychopathological phenomena specifically related to time like
chronophobia, chronotaraxis are sometimes stressed, and correlated with
localizable cerebral lesions. The time threshold in the optical area is judiciously
referred to as the flicker-fusion frequency, the extreme case of fusion occurring
in our perception of colors, which bears no witness to the undulatory nature of
the electromagnetic stimulus. The shortest per-ceived duration has been
determined and called a psychological moment: a human lifetime is estimated as
being of the order of 20 billion moments. The problem of time symmetry and
reversibility has been discussed in a cursory way. We are told that "biologists
learn that there must be an irreversibility of the course of time because men get
old and things disintegrate". The planetary motions, admittedly run by time-
reversible laws, are yet detennined by a sun which is getting old. In the
astronomical area, things disintegrate but astronomers are now talking of a time
symmetrical cosmology. The fact that the single most significant phenomenon in
biological evolution, viz. mutation ala de Vries, is time reversible, is hardly
noticed.
Interesting results have been obtained in the cerebral localization of man's
intuitive ability to estimate the length of time intervals. According to
H. Hoagland 103 and H. von Foerster,104 the human time sense depends
primarily upon the velocity of the oxidative metabolism in the cells of the brain,
perhaps of the reticular formation in the brain stem. The Arrhenius
180 DURATION AND CAUSALITY

equation which connects the speed of the oxidation with the activation energy of
the relevant neurons (where activating energy is construed as the amount of
kinetic energy per mole which the molecules must acquire before they react) was
successfully used by von Foerster and Hoagland. Incidentally, this is a most
significant corroboration of the principle of psychophysical parallelism
(discussed in Chapter 3 of this part). Parallelism was required to establish the
possibility of consistently fitting the physical, psychological and psychophysical
temporal orders into a single cosmic time. Parallelism requires that every mental
experience have a cerebral or endocrine substratum. The aforementioned
scientists show that the existence of a cerebral substratum can be established for
the most crucial mental experience, viz. that of tem-poral interval.

6. THE EPISTEMOLOGICAL SIGNIFICANCE OF INTUITIVE TIME

One is no doubt tempted to view the intuitive experience of time, especially the
perception of those temporal figures where the structure of time becomes almost
palpable, as a direct and representative apperance of infmite and universal
duration. This thought may make two forms. In its ontological form it consists
in the belief that the true nature of the temporal order manifests itself only under
the exceptionally favorable conditions presented by intuitive experience, in
which the object of the cognitive act is an almost simultaneous attachment of the
self accomplishing this act, so that the gap between the knowledge and the thing
known is reduced to a minimum. Science, with its many devious approaches to
the temporal order, only knows it indirectly and under artificial and distorted
forms. The true nature of time would therefore be just as it appears under the
favorable proximity found in intuition, in which the relations of before, after and
during, conceived as engendering a stratification and an essential orientation of
cosmic becoming, would be only
a continuation of the order of succession and simultaneity intuitively given in
the infmitesimal region accessible to direct experience. The consideration of
visual sensations will best serve to bring out the hypothetical and, after all,
gratuitous character of this extrapolation of the intuitive aspect of time. The
perceived colors are entities whose nature seems to be fully revealed by their
sensory aspect. Their resemblances and contrasts, their dazzling changes and
their almost solid fixity seem to impose themselves on the mind with perfect
clarity, and nothing is more natural than to suppose that the true nature of the
physical process is clearly revealed here, and that this nature remains the same in
the analogous physical processes which, known under unfavorable
INTUITIVE FOUNDATIONS 181

conditions, no longer appear as colors. But this has not been the solution
pwposed by science. Along with the electromagnetic vibrations corresponding to
our sensations of color, it readily admits vibrations which do not manifest
themselves visually, and believes that visual data make up only an infinitesimal
segment of the whole range of electromagnetic vibrations,just as the temporal
interval accessible to intuitive experience constitutes only an infinitesimal
fragment of universal time. The scientist does not try to attribute sensible
qualities to invisible waves, and views the qualities associated with visible waves
as a human adjunct with no intrinsic physical significance. Why should the
intuitive aspect of time be any more privileged? In our opinion, this aspect has
only limited significance for the fragments of inner becoming; the temporal order
established by scientific methods, whether it concerns psycho-logical, inter-
psychological, psychophysical, or physical time, has throughout a causal
character.
The epistemologjcal interpretation which can be given to the extrapolation
of the intuitive temporal order is more important. It amounts to saying that the
sense of before, after, and during applied by the scientists in all the sciences of
the real is precisely that of intuitive simultaneity and succession, since it is in
intuitive experience that we learn what it means for a pheno-menon to take place
before, after, or Simultaneously with another. The visual example is more
favorable here. We would say: 'The term 'red' is comprehen-sible only to someone
who has seen at least one red object, and it is from this experience that he would
have learned the meaning of the term. He may certainly then attribute the color
red to an object not directly given or even inaccessible to direct experience either
because of its location or because it is too large or too small. But in attributing
this color to the unperceived object he uses the term in a sense deriving from a
situation where the object was directly given. A man born blind would not be able
to make this judgment, since he would not understand it at all'. I believe that this
reasoning fails to take into account the distinction between a sensory object and a
perceived object. It is in its application to a sensory object, to a colored patch in a
subjective visual field, that the term 'red' has an intuitive meaning. Applied to a
perceived object situated in physical space this term changes in meaning, to
approximately this: 'capable of appearing red to a normal observer, suitably
located, and in daylight.' This latter is therefore comprehensible even to a
congenitally blind man, who, using indirect tactile methods, will be able to decide
whether a particular object has the property of provoking in a given observer in a
given situation a sensation which this observer will call 'red'. Therefore, it is not
the intuitive meaning which is extended to cases
182 DURATION AND CAUSALITY

inaccessible to direct experience, but a new meaning related only indirectly


to the intuitive one. I believe that the same is true of universal succession
and universal simultaneity. The definitions of the different components of
uni-versal time will illustrate the fact that it is not synonymous with intuitive
time, and consequently, that the epistemological extrapolation of intuitive
time is inadmissible.
Notice that even psychological time cannot be identified with intuitive time. It
follows directly from the existence of gaps of the second kind that there are
psychological states which belong to the same self, and consequently to the same
psychological time, but which nevertheless bear no intuitive temporal relation to
each other. In order to locate them in time we are forced to employ the same
indirect methods which are used to synchronize psycho-logical states belonging
to different fluxes. Furthermore, these methods are practically inevitable in the
majority of the cases where intuitive time (Le., intuitively knowable and not
intuitively known) displays no gaps. It is only a more or less limited portion of
our past, which could be called the living past, whose states we are capable of
reliving and whose temporal structure we can therefore penetrate. Behind it
stretches a temporal region which we very abstractly recognize as having been
our own, but which no longer forms a living whole with our present. Our
childhood photographs, which we look at with incredulity, our writings, which
we reread after many years, finding ourselves unable to understand them
completely or to re-immerse ourselves in their atmosphere, all are related to our
self, but the past which they reflect is psychologically dead. It is not in direct
memory that we view these no longer familiar shadows of our past, but through
all kinds of material inter-mediaries, just as the psychological states of our
neighbor are in general accessible to us only through the bodily gestures which
express them.
Thus, to fill the gaps in our knowledge of the temporal structure of our
past it seems necessary to go beyond the framework of psychological data.
This can be done in two ways. The complete picture of the universe exterior
to a given person comprises the psychological states of the people around
him, as well as the physical events which take place in his surroundings.
From the epistemological point of view, whether to extend intuitive time by
means of the physical data or by means of the inter-psychological data seems
arbitrary up to a point. One can, without leaving the realm of solipsism,
group connected images into systems corresponding to 'physical things' and,
in particular, construct a material clock whose indications would serve to fill
the gaps in intuitive time. It will undoubtedly be objected that this possibility
of a solipSistic physics is entirely fictitious, and that physics, in its historical
INTUITIVE FOUNDATIONS 183

evolution, is an eminently social product; in its realization, beyond the genius,


however great, of a single individual. But the solipsist is by no means forced to
refuse the collaboration of his colleagues, especially not the information they will
give him about the physical universe, which will complement his knowledge of it
in an essential way: he will simply treat his colleagues as very precious recording
or calculating instruments and will conSider, for example, an oral report by one
of them as a train of sound waves which is extremely instructive with regard to
the spatio-temporal environment of the body emitting these waves. By thus
grouping sensory images into systems meant to correspond to physical things and
by complementing these data with those furnished by the various recording
apparatus (including live apparatus) he will arrive at a picture of the physical
universe which is no different from that of science, without making any
hypotheses about the psychological life of others like him. It will be only after
having drawn this picture of the physical universe that he will decide to complete
it with data about the psychological life of others, and examine, in particular, the
temporal structure of this life.
But this procedure can also be reversed: one may complement the intuitive
data with hypotheses about the psychological life of others, without grouping
sensory images into systems corresponding to physical things. In perceiving
someone's face, I can physically objectify the perceived image, by including it in
a system of images all of which belong to this face, but I can also objectify it
psychologically, by relating it to a psychological state of this same person. There
seems to be no essential epistemological or psychological difference between the
perception of physical things and the perception of the psy-chological state of
others. It is neither the thing nor the psychological state that is immediately
given, but rather a sensory image, which is interpreted sometimes as an aspect
of. the thing, sometimes as an expression of the psychological state. In both
cases, this interpretation is not an implicit conclusion drawn from the description
of the image, but rather an awareness of the presence of the thing or person in
question, which is of dubious epistemological value. To justify hypotheses about
the existence of physical things, we deduce from them verifiable consequences
dealing only with sensory data. The same procedure is applicable to hypotheses
concerning the psychological life of others. One who wished to formulate the
latter independently of the former might say, for example: "Among the images
appearing in my visual field, I can associate some with my body by virtue of
some unique properties which can serve to defme them (thus, the image of the
surface of my body has the unique property of being followed by a double tactile
sensation after being approached by the image of my finger, 105
184 DURATION AND CAUSALITY

while similar images of the bodies of others are not associated with these
double sensations). By perceiving then that certain images of my body are
associated with certain of my psychological states (e.g., the image of tears
with pain), I will be able to interpret the image of tears on another face as
corresponding to a pain which is not mine. By definition, this pain will be
simultaneous with the image I have of the corresponding tears."
We see, then, that the theory of inter-psychological time does not essen-tially
presuppose that of physical time and that, consequently, one can expound
physical time before or after inter-psychological time without distorting the ideal
succession of phases in the epistemological genesis of time. The same is true of
psychophysical time. When a friend smiles at me on the street, I am convinced of
intuitively grasping the simultaneity of his pleasure both with mine and with his
gesture of greeting. Neither of these two relations of simultaneity - inter-
psychological and psychophysical - implied in this situation, seems to be
epistemologically prior to the other. However, we will see later that it is only in
cases where the psychological life of another seems to be directly given that
extra-physical time is epistemologically co-ordinated with physical time. In every
other case, we arrive at a hypothetical knowledge of extra-physical time with the
aid of hypotheses about physical reality. This is why, in the present part of this
book, we will place the causal study of physical time before that of extra-
physical time.
Another way of expressing this ideal possibility of extending intuitive time
either on the physical or the inter-psychological side would be to say that
physical time and interpsychological time can both become intuitive, and thus
form no epistemological hierarchy. Obviously, it is only in a derivative sense
(although an important and even inevitable one from a certain point of view,
which we will indicate) that one can speak of an intuitive physical or inter-
psychological time, since, strictly speaking, only psychological time is capable of
being known directly and indubitably. Thus, for example, the temporal order of
physical events can be considered intuitive insofar as it can be identified with the
intuitive order of their perceptions. This is not always the case; the successive
perception of lightning and thunder in no way involve the non-simultaneity of
these physical events. But it is clear that physical time would not be knowable at
all if it were not intuitively knowable, in at least some cases, i.e., if the temporal
order of physical events were not in certain cases the copy of the order of their
perceptions. The scientist really thinks that he knows the temporal order of the
paleontological epochs, although no man has witnessed them. But to establish
hypotheses relating to them he uses fossils, skeletons, and clocks, whose
positions and changes are
INTUITIVE FOUNDATIONS 185

intuitively ascertained. If he did not believe that simultaneously perceived parts


of the skeleton were simultaneous, that successively perceived positions of the
hands on the clock did in fact succeed one another in time, he would be able to
imagine neither the paleontological facts nor their temporal rela-tions. It is
therefore beyond question that, within certain limits (viz., con-cerning local
perceptions of events near the observer in space and time, and involving neither
high velocities nor extreme dimensions), the scientist is forced by the very nature
of his investigations to consider the temporal order of physical events as a copy
of the order of their perceptions, and that it is only the extension of the temporal
order beyond these limits which necessitates the causal methods whose axiomatic
foundations will be studied in the following chapter. The intuitive knowledge of
time is at the very basis ofthe knowledge of physical time, just as well as of
psychological time.
Must we conclude that the statements concerning the fundamental proper-ties
of intuitive time are a priori principles, irrefutable by any experience, since it is
only with the aid of these principles that we can establish any experience? Or
must we consider beliefs about the properties of this time as hypotheses which,
up to the present, have been verified in an extremely large number of cases and
which consequently seem practically unquestion-able? I believe that experiences
refuting the laws of intuitive time are neither unimaginable, nor perhaps,
unrealizable. Let us consider, for example, the connection between perceived
time and the time of memory. We have seen how the sensory temporal order,
perceived directly at the heart of the ex-perienced present, is extended further and
further by the function of memory, which makes us aware of the perpetual sliding
of the present into the past. To the succession directly given by memory, there
corresponds a sensory and isotropic succession. What guarantee have we that
these two temporal orders, which we always find intimately linked to each other
in our own past, cannot sometimes become dissociated, that the temporal order of
recollections, instead of remaining identical to the sensory temporal order, cannot
defme a new temporal dimension in the ensemble of our psychological states? We
can well imagine the psychological life of a man each of whose external
perceptions would be followed by an infinite series of internal perceptions, the
temporal order of the external ones being based only upon the inter-connection of
memories, and the temporal order of the internal ones being sensory. The intuitive
time of this man would form a two-dimensional con-tinuum. The dimension of
memories would have transsubjective significance and would synchronize his
psychological life with universal becoming, while the sensory dimension would
be exclusively subjective and would bear no
186 DURATION AND CAUSALITY

intrinsic relation to cosmic time. But we can also imagine that it is the dimen-
sion of sensory time which corresponds to universal time, while the dimension of
memories has only subjective significance. A man whose intuitive time
underwent a dissociation of the first kind would have a very good memory of
each instant in his past (in the objective sense of the term), but his present would
decompose itself into an infinite series of psychological states having the same
structure as a very short segment of a normal conscious flux. A man whose
intuitive time underwent a dissociation of the second kind would constantly
forget his past (in the objective sense of the term), but would know his entire
present with the aid of memory. A dissociation of this kind is brought to mind by
the descriptions of the psychological states provoked by the action of certain
narcotics, where the individual seems to live subjec-tively through a period
which is extremely long in comparison with the time which actually elapses. The
very possibility of these dissociations proves that the union of the isotropic time
of perception with the anisotropic time of memory, which constitutes the very
nature of intuitive time, has only an empirical character. The same is true of the
other properties of this time, e.g., the asymmetry of succession. In considering
defmition (I), we see that this asymmetry rests essentially on the fact that if
psychological state X contains a recollection of states Y, Y will not contain in turn
a recollection of state X. This is unquestionably verified by innumerable
observations, but it does not seem impossible that it should fail in some particular
case, that a particular individual should pass through a psychological state.
containing the memory of a state which succeeded it from the point of view of
universal time. If serious observations were to lead us to admit this fact, we
should not reject it in the name of some a priori necessities, but we would simply
conclude that in this individual, intuitive time had taken a very peculiar form.
IX

PHYSICAL TIME

1. GENERAL REMARKS ON THE CAUSAL THEORY OF PHYSICAL TIME

Intuitive time has been defined as a set of relations of simultaneity,


succession, and duration linking psychological states to one another. It is
neither a homogeneous medium where psychological processes follow their
course, nor a receptacle in which they take place; it has no reality outside of
them. In rejecting the substantival theory of intuitive time we made no
dogmatic assertions on the nature of time, but only contended that there is
nothing in what we learn from observation of the temporal properties of the
psycho-logical flux which forces us to accept the substantival nature of time.
Can these considerations be extended to physical time? Can we reject the
hypothesis of a substantival physical time merely by invoking the argument that
such a time is foreign to every direct experience, that the only perceptual data
concerning physical time have to do with the simultaneity, succession, and
duration of physical events, and that the addition of a class of instants, considered
as fundamental and irreducible entities, to that of events is a gratuitous
hypothesis? We think not. If physical time is not directly per-ceivable, neither are
physical events. The identification of physical objects with homonymous
perceivable objects cannot be maintained. Noises, colored surfaces, motions, as
perceivable objects endowed with the sensory qualities revealed in their
perceptions, cannot be identical to the mechanical and electromagnetic processes
which the physicist coordinates to them, since he attributes to these processes a
large number of properties which are incompatible with those appearing in
perception. And it makes little difference here whether we consider the physical
universe to be a set of constructed or inferred entities, or whether we interpret the
propositions of physics as hypotheses concerning the real or as conventions
creating a fictitious universe. What is beyond question is that the basic difference
between the physical universe and the sensory universe makes it impossible to
interpret the corre-spondence established by the physicist between the sensory
datum and the physical process as implying their numerical identity. Thus, the
fact that sub-stantival physical time is never perceived by no means constitutes
sufficient
187
188 DURATION AND CAUSALITY

grounds for rejecting the hypothesis of its existence; the same argument
would lead us to reject any other physical hypothesis.
There are, however, other arguments which seem to us to be decisive. The
principal fact upon which the relational theory of time rests is that a class of
instantaneous events contemporaneous with a given instantaneous event is
endowed with the same ordinal properties as are attributed to the instants,
and may therefore serve to define them. Yet, it is not this defmition of the
instant which constitutes the relational theory, since nothing in the sub stan -
tival theory prevents us from defining classes of.events simultaneous with a
given event and noting that these pseudo-instants enjoy the same linear order
as the true instants. The trick of the relational theory consists in considering
only these pseudo-instants of the substantival theory, ignoring completely
substantival hypotheses about the instants themselves. It therefore employs
fewer hypotheses than the substantialist theory, since it admits only the
existence of the pseudo-instants, while the substantival theory adds to these
hypotheses about the existence of the instants. From this point of view, the
relational theory seems distinctly preferable.
This methodological argument is applicable to the problem of absolute
time, as well as to that of relative time. But the question takes on a new
aspect if we take into account the relativity of time; the hypothesis of a
unique time composed of instants arranged in a linear order then seems
unintelligible. It is true that certain authors, while rejecting absolute time,
continue to speak of instants, attributing to them a 'conic' instead of a linear
order .106 But this is clearly only a terminological question. The instants
partaking of the conic order are infmitesimal regions of space-time - world
points; to affIrm that they are elements of time amounts to identifying time
with all of space-time, which is quite different from asserting the sub-
stantiality of physical time.
The relational theory of physical time seems therefore to be more in
keeping with the present state of science. Granted, it does not provide us
with very much information about time: to know that time is an order and a
magnitude, and not a substance, does not tell us very much about this order
and this magnitude. But it points the way to a solution of the problem; it tells
us that the true problem of time resides in the analysis of temporal relations.
Let us emphasize the scope of this problem, which has been obscured by
recent forms of the relational theory. Leibniz's theory seems to be generally
accepted today, but the positivistic interpretation put upon it unduly restricts
the scope of the problem of time. Thus, very often we identify physical time
with the set of readings of clocks; the properties of physical time would
PHYSICAL TIME 189

concern only the behavior of the instruments which serve to measure it. To us
this seems to be a confusion of the measure with the thing measured. Admittedly,
the temporal structure of becoming influences the behavior of clocks. But the
converse is not true. Becoming would possess temporal properties even if no one
had ever thought of measuring them with a clock.
However, this positivistic interpretation of the relational theory of time is
attenuated in practice, since one extends the concept of a clock to every physical
phenomenon capable of serving to measure time, in particular, to the behavior of
light and a rigid body (the face of the ideal clock, described by Einstein and
Carath6odory, 107 is made precisely with rigid rods, and its hands are light rays).
Thus, the positivistic interpretation is found to be quite expanded, since it
amounts to saying that the properties of time are essentially the properties of light
and the rigid body. Furthermore, we might add to this the particle (material point)
moving without the influence of ex-ternal forces, which, according to Weyl, lends
itself very well to the definition of the metric of space-time. We might even add
the point moving under the influence of a field of force, since, as Infeld has
shown, this point could also be used in the construction of an ideal clock. We can
see that these successive extensions of the notion of clock deprive the positivistic
interpretation of the relational theory of any precise meaning, since the properties
of measuring instruments, which it takes to be the properties of time, become
almost any property of almost any physical object. It might be said that these
extensions of the notion of clock are purely fictitious, and that the important thing
is to stick to those measuring instruments upon which the scientific knowledge of
time really rests. But science unquestionably builds its clocks from rigid rods
and employs electromagnetic signals to synchronize them. Must we not conclude
from this that all we really know about the temporal structure of physical
becoming concerns, in the final analysis, the properties of light and the rigid
rods?

I do not think so. Reichenbach tried to show that we can derive the properties
of space-time from certain postulates concerning the propagation of light,
independently of any hypothesis about the properties of matter [provided that
some additional assumptions, overlooked by him and related to Bateman's
theorem be made. Cf. Part III below. H. M.] But, on the other hand, Infeld has
attempted to prove that we can derive the Lorentz formulas (which summarize
the structure of space-time) by making certain simple hypotheses about the rigid
body analogous to those of the theory of elasticity. But doesn't the possibility of
these two constructions, one of which is based exclusively on the properties of
light and the other on those of the rigid body,
190 DURATION AND CAUSALITY

prove that the spatio-temporal structure of the physical universe is reducible


neither to the behavior of light nor to that of the rigid body, and that,
consequently, it must rest on a more fundamental property of becoming,
capable of subsisting in regions devoid of light and rigid bodies? Moreover,
it is easy to indicate physical systems whose spatio-temporal structure is
determined independently of light and the rigid body. For example, this
would be true of a system where the measure of time would be defined by
the disintegration of a radioactive element (with negligible fluctuations), and
spatial length by that of an inelastic thread. In this system there would be no
fixed coordinate axes, and spatio-temporal measurements would have to be
made by methods other than those which take a rigid trihedron as a point of
departure, but this difficulty is not insurmountable. A less artificial example
is that of the bottom of the ocean. There, light scarcely penetrates. Under the
enormous pressure which is exerted there, solid bodies, in general, do not
behave like rigid bodies. Yet, no physicist would say that because of this
absence of light signals and rigid rods at the bottom of the ocean, the spatio-
temporal structure of the events which take place there is any different from
that at the surface of the globe. [The center of the sun, with a temperature of
the order of twenty million °C, yields a more telling example. H. M.]
But perhaps a supporter of the positivistic interpretation of the relational
theory of time will not be convinced by these examples. He will object that
the belief in the normal properties of the spatio-temporal order of events at
the bottom of the ocean is unfounded, since there is nothing to guarantee that
these properties do not change under high pressures, and that it would be
necessary to go down to the bottom of the ocean with appropriate measuring
instruments in order to make certain. Until such measurements were made,
these properties would remain unknown. And similarly, he will object that
the difference between the optical and mechanical constructions of Reich-
enbach and Infeld on the one hand, and a more general concept applicable a
priori to regions without light and rigid bodies on the other, is only apparent,
since every case which verifies the latter also verifies the former. Two theories
differing only in their unverified consequences would for him be epistemo-
logically equivalent.
It is true that, as it is usually presented, the optical method implies the fiction
of messengers travelling everywhere and always (or the fiction of clocks filling
the universe), while the mechanical method is forced to fill all of space with
myriads of rigid trihedrons which miraculously avoid one another in their
uniform, rectilinear movements. But the advocate of the positivist interpretation
could modify this method so that it had no more
PHYSICAL TIME 191

fictitious components. This modification would consist in considering the optical


and mechanical criteria of the temporal order as partial, not total definitions.

One example will be enough to show the difference between a total defini-
tion and a partial one. Equality of weight is said to be defined physically by a
pair of scales: two bodies are equal in weight, if, when placed on either side of
an accurate scale, they remain in equilibrium. Obviously this condition of
equilibrium, which is sufficient for determining the equality of weights, cannot
be realized where two bodies are not on a scale. Can we say that in this case it is
at least possible to place the two bodies on an accurate scale while maintaining
the equilibrium of both sides?
We have seen, however, that this notion of possibility is completely im-
precise. IOS We cannot accept the fact that once the two bodies are removed from
the scales they cease to be equal in weight since this would contradict precise
physical laws. We would avoid this conclusion by considering the equality of
weight as undefined for these two bodies, and its definition just cited as a partial
defmition.
The same consideration would seem applicable to optical and mechanical
criteria of the temporal order, for example, to the optical criterion of simul-
taneity, regarded as a partial defmition of simultaneity: simultaneity would be
defmed only for pairs of events linked by light messages (or better, for events
near clocks synchronised by light messages). The other pairs of events would not
be non-simultaneous, but rather void of any defmed temporal relation. Thus
conceived, the positivistic interpretation of the relational theory of time
undoubtedly avoids the above-mentioned difficulties. But an examination of the
partial defmitions it uses leads to many other difficulties. Notice, first of all, that
these definitions, as propositions, are doubly hypo-thetical in form. Their general
scheme is this: if object 0 has property ~, in order for 0 to have property 71, it is
necessary and sufficient that it have property ~ (1).109 71 is the defined property,
~ the defining property, while
~ designates the condition of the definition's applicability (for the equality of
weights, it is the fact these are on an accurate scale; for simultaneity, it is the fact
of being linked by light messages). This scheme shows that of two partial
definitions of a notion, the preferable one is one whose condition of applic-
ability is most general, and that the partial definition becomes total when its
condition of applicability is identically filled by every object.
Now, it is certainly possible to designate the propositions of form (I) as partial
definitions, even if their conditions of applicability are very res-tricted. But this
does not relieve us of the task of searching for more general
192 DURATION AND CAUSALITY

conditions of applicability and above all for a total definition since current
terminology by no means admits that the weight of a body is undefined the
moment it is off the scales and that, similarly, the simultaneity of two events not
linked by a light message is in no way considered undefined. Basically, if the use
of a term rested only on a partial definition, for example, if this term were
introduced into the language by means of this definition, this would amount to
considering this term as undefinable and its partial definition as an axiom which
it satisfies. The only additional postulate which could be considered as
characteristic of partial definitions is that this axiom lends itself easily to
verification, that is, its condition of applicability is easily realizable. But this
postulate hardly excludes conditions of applicability which are too restricted.

[This form of reductionist interpretation, however, would unjustifiably curtail


science. There may be good reasons for putting scientific theories on a par with
the same verifiable consequences, but to speak in terms of verified
consequences only, would shrink the whole scientific world-picture to a hodge-
podge of events which happened to be witnessed by some scientist. Even if every
temporal hypothesis could be verified by using only optical methods, it would
still be true and important that such hypotheses could also be verified by
examining, for example, their mechanical consequences. The latter are therefore
as integral to the relativistic view of time as the former.
One may doubt, moreover, whether the entire relativistic structure of space
and time can be defined in terms of laws governing the behavior of light signals.
To establish the possibility of such a definition it would be necessary to show
that only the Lorentz group of spatio-temporal transformations is compatible
with these laws. Yet, according to a remark made by H. Weyl, the laws referring
to light signals (including the constancy of the velocity of light), do not single out
the Lorentz group, but rather a wider group obtained by superimposing arbitrary
similarity transformations upon those of Lorentz. 11 0 This means that light would
behave intrinsically as it does even if the dimensions of material objects (such as
rigid yardsticks) and the tempo of material processes (e.g. those occurring in
material clocks) were distorted beyond recognition. Axiomatizations of the
relativistic space-time structure based upon the properties of light signals (e.g.
those of CaratModory and Reichenbach) are therefore open to a serious
objection'which does not arise if the relativistic theory is not restricted to optical
phenomena.
Let us add that in order to realize this reductionist objective, the relativistic
theory must not be restricted to optical phenomena. To the behavior of light
signals (or, alternatively, to the properties of rigid bodies as exemplified in
PHYSICAL TIME 193

yardsticks and clocks), the reductionist axiomatization must add the fiction
of light messages travelling everywhere and always, or fill the whole
universe with clocks and yardsticks moving in every direction with every
possible speed and yet adroitly avoiding one another in their peregrinations:
such clocks, yardsticks, and light signals have to be omnipresent for their
structure to be coextensive with the pervasive spatio-temporal medium of
physical reality. Fictions are certainly no novelty in the history of science,
and Vaihinger's philosophy of science is by no means obsolete. In the most
fundamental of the sciences - physics - these fictions, called 'idealizations',
are an essential methodological device. Yet, in all important cases, science
has come sooner or later to eliminate the crucial fictions or assign a merely
auxiliary role to them, e.g., that of abbreviatory devices which could be
dispensed with in principle. Thus the ideal thermal machines with infmite
heat reservoirs which dominated the classical period of phenomenological
thermodynamics have eventually been shown by Caratheodory to be 111
replaceable with unobjectionable mathematical concepts. The fictitious
'delta-functions' which formed the mathematical core of Dirac's codification
of quantum mechanics 112 have been shown by von Neumann 113 to be
dispensable. The aim ofaxiomatizing the relativistic space-time theory is to
exhibit its fundamental assumptions and to provide an insight into its internal
logical structure and its observational basis. This objective is hardly secured
by imputing a fictional world view to the theory. Such a view would remain
unsatisfactory even if it helped to make time less puzzling and elusive by
interpreting the relativistic theory of time in terms of such familiar entities as
light signals, clocks, and yardsticks.]
Thus, neither the properties of light nor those of rigid bodies provide
conditions of applicability sufficiently universal to establish the structure of
physical time. But each of these groups contains this structure, and this
indicates the direction in which one must look for more fundamental proper-
ties, which would be sufficient and indispensable for the temporal order.
What, essentially, is a light ray? It is a group of events linked by the causal
relation and endowed with certain additional well-defmed properties. What is
the history of a rigid body? It is again a group of events linked by the causal
relation and endowed with certain other additional well-defmed properties.
This suggests that the more general notion which should serve as a basis for
the structure of time is that of a group of events linked by causality, or - what
amounts to the same thing - that of the causal relation.
Certain simpler and more direct considerations lead to the same result. Let us
imagine two regions of the universe (such as the New and the Old Worlds),
194 DURATION AND CAUSALITY

which, during a given period of time, do not act upon each another, since
they were separated at the beginning of this period and joined only at its end.
The history of each will be discoverable to any desired degree of precision
by an investigator who, after the regions are reunited, questions witnesses
who had lived in both of them during the period of separation. However, this
investigator, if asked to synchronize the two histories, to say what was taking
place in the first region at the time when such and such an event took place
in the second, would be unable to reply; if he follows this analysis through
he will be forced to deny any precise meaning to this question. It would have
been otherwise if, during the period in question, there had been an exchange
of causal actions between the two regions. These actions wouldoereate a
com-mon time, in which both histories would be jointly located. It is not
necessary that there should always have been signals, messages traveling
between the two regions; it need only be possible to form a 'causal
decomposition' 114 with the set of events constituting the two histories which
would be the case if each of them contained events produced under the
influence of the other; it would then be this influence which would create a
time common to the two histories.
More simply, let us imagine that two physicists working in adjacent
rooms had observed respectively the motion of a strong magnet and the
displacement of iron filings along the partition separating the two rooms.
Thanks to the causal relation between the motions of the magnet and the
filings, the two physicists, conferring together afterwards, would be able to
synchronize the events which took place during their respective stays in the
two rooms, even if they had not been able to consult previously
synchronized watches. We are pointing here to the greatest generality of the
method of decompositions with respect to that of messages: it makes
possible the synchronization of the histories of the two rooms, even though
no message has linked them during the period of time in question.
One might object that the observers respectively situated in the two
separated regions will not be forced to look for causal interaction between
the regions in order to synchronize them, since it would suffice to compare
the histories of their respective regions with that of some common pheno-
menon: the two physicists need only observe the sun through the windows of
their rooms, and, basically, the same procedure would serve to synchronize
the histories of Europe and America during the the period of their separation.
This objection 115 duplicates a thought of Kant's - it represents one of the
causal defmitions of simultaneity appearing in the Critique of Pure Reason.
Yet it is clear that the use of the solar clock suppresses the causal separation
PHYSICAL TIME 195

between the two regions. These two, along with the sun, fonn a system where a
unique time is defined by means of the interaction of its parts.
This thought, that the time of things is inseparable from their interaction, that
it expresses their causal interdependence, constitutes the principal theme of the
causal theory of time, and is therefore far from being paradoxical. Let us recall
briefly the objective of the causal theory of physical time, such as it emerges
from our critical studies in Part I of this book. The aim of this theory is not to
make intuitive knowledge of physical time useless but to extend the laws of time
furnished by this intuition to regions inaccessible to it. The causal theory does
not replace, but indirectly generalizes 116 the intuitive data of common sense, and
it must coincide with these wherever they are applicable. However, this
generalization deprives physical time of several characteristics essential to
intuitive time: the causal order is relative and isotropic, while the intuitive order
is invariant and anisotropic. The discovery of the relativity of physical time, no
matter how conventional the fonn of its definition, has seemed to us to be that of
a new causal structure of becoming. The anisotropy of time has appeared to us, in
spite of the efforts of its defenders, to be a local property, derived from the
history of the material universe rather than from its physical theory. In order to
establish a theory of this isotropic and relative time, independently of the
complicated and often inapplicable notion of inertial system, we have proposed a
'method of causal decompositions' whose aim is to determine the place in the
spatio-temporal order occupied by each event in terms of the relations of
causality which link this event to the set of other events. It is to the systematic
devel-opment of this method, to its axiomatization, that we shall proceed in this
chapter.

The causal theory of time to which we are led by the method of decom-
positions differs from the theories we have summarized in the first part of this
book in that it takes symmetrical causality as the only supposedly, undefined
notion. We will adopt this procedure in what follows because it allows us to
deduce the ordinal and quantitative properties of time from a few simple
properties of the causal relation, and also because of certain difficulties inherent
in the causal theory of time which it enables us to overcome:

(a) As usually understood, the relation of cause to effect is asymmetrical, the


cause, being always earlier than the effect, is consequently incapable of
becoming the effect of its effect. It is therefore the asymmetry of before and
after which is found in the asymmetry of causality. This is why, in our opinion,
the explanation of before and after in terms of the relation of cause
196 DURATION AND CAUSALITY

and effect involves a vicious circle, the fundamental difference between the
cause and the effect residing precisely in the priority of the cause with
respect to the effect. Furthermore, we have seen that the attempts made to
distinguish cause from effect by using extratemporal criteria are open to
serious objec-tions. ll 7 This difficulty vanishes if, instead of considering this
asymmetrical relation of cause and effect, we consider the symmetrical one,
which indiffer-ently relates cause to effect and effect to cause.
(b) It often happens in practice that we are sure of the existence of a causal
relation between two events without being able to say which of the two is the
cause of the other. 118 When a container of gasoline has burned during the
burning of a house, we do not doubt that there is a causal relation between these
two events, often without being able to distinguish which is the cause of the
other. But the converse is not true. In being able to distin-guish cause from effect
in a pair of events, we thereby ascertain the existence of a causal relation
between them. By basing the theory on symmetrical causality, we formulate more
restricted hypotheses, thereby assuring our theory of a richer contact with
experience. In particular, it will contain as special cases the axioms based on the
use of light signals and rigid frames of reference, and will be verified in all cases
where these axiomatic systems are verified.

(c) Finally, there is the problem of the reversibility of physical phenomena


which we discussed in detail in the first part of this Essay. If we adopt asym-
metrical causality, we would thereby also adopt the point of view of irreversi-
bility which is by no means rendered probable by the present state of science. The
use of symmetrical causality allows us to treat the theory of time in-dependently
of the problem of reversibility; it is possible to add the axiom of irreversibility to
our theory, but this axiom is independent of the others and is superfluous to the
foundations of the theory of time. Furthermore, this procedure seems to be
advantageous even from the point of view of irrever-sibility. Eddington, who
believes in irreversibility, and deduces from it an anisotropy of physical time,
nevertheless points out that there could be regions in the universe where a
statistical eqUilibrium is attained, and con-sequently where, entropy having
reached a maximal and stationary value, nothing would permit the attribution of
any privileged direction to physical time. But time would not disappear from
these regions; its topology and
metric would retain their essential properties. Consequently these properties
must be independent of the problem of reversibility. It is this independence
which we wish to bring out by basing our theories of time on symmetrical
causality.
PHYSICAL TIME 197

2. SYMMETRICAL CAUSALITY

It goes without saying that in the axiomatization of a physical theory the choice
of primitives is of primary importance. These terms guarantee empirical meaning
to the deductive theory, which, taken by itself, could be applicable to all kinds of
abstract entities, as well as to the real order of phenomena. We will therefore
begin with some comments on the notion of the causal relation, which
supposedly is the only indefinable notion in the method of decom-positions.
These preliminary remarks must form a commentary on the notion of this causal
relation and thereby suggest suitable axioms. But they are by no means part of
the axiomatic system, and the definition of the causal relation to which they lead
is not utilized in this system. 119
The causal relation between two events 120 expresses the existence of a
necessary connection between them. This link can relate only events of a certain
spatio-temporal orientation, and this is precisely why the structure of space-time
is defmable in terms of the causal relation. Just what is this necessary
connection? Even since Hume, it has been explained as the regular succession of
two events: the reproduction of the cause involves that of the effect. In this
formulation, it is the term 'reproduction' which is not very clear. It goes without
saying that it is not an integral repetition: the cause, as an individual event,
cannot be repeated with all of its properties, since, according to the principle of
the identity of indiscernibles, the community of all the properties would involve
not only the specific but also the numeri-cal identity of the two cases, and would
render the repetition impossible. Therefore, in this idea of reproducibility of
physical phenomena, we are dealing with a partial community of properties.
Which properties are re-peated? How closely must two physical events resemble
each other for one to pass as the reproduction of the other?

Posed in this abstract way, the question seems very vague, but this is no
longer so if we take into account the concrete form of the laws of physics.
Thanks to the extraordinary unification of physical theories resulting from a long
and fruitful evolution, the fundamental properties whose causal inter-connection
is studied by the physicist, and which are completely sufficient for the
characterization of the most diverse events (such as those in the examples cited
above), have become very few and can be simply enumerated. The spatio-
temporal distribution of masses, of the intensities of electromag-netic and
gravitational fields, and of the eigenvalues of occupation-number operators
which correspond to elementary particles known to exist, are the only properties
of events considered by the physicist when he speaks of
198 DURATION AND CAUSALITY

reproducible events: for him, two events are specifically identical if they
exhibit the same distribution of masses, field intensities and occupation-number
operator eigenvalues. In calling these fundamental properties studied by the
physicist 121 'intrinsic' we should state the definition of the causal rela-tion
among physical events in the following manner: event A, followed by event B, is
said to be the cause of B if each event A', having the same intrinsic properties as
event A, is followed by an event B' intrinsically identical to B
(1). Thus, the proposition asserting that the passage of a light ray from one
optical medium to another causes its refraction would mean that every event
characterized by the same distribution of masses, field intensities and eigen-
values of occupation-number operators as the event of the passage of the light
ray from one medium to the other, is followed by an event exhibiting the same
distribution of these magnitudes as the event of the refraction, i.e., intrinsically
identical to the latter .122
This explanation, defining the causal relation in terms of reproducibility and
succession in time, hardly mentions the spatial factor, which nevertheless seems
essential for physical causality. Let us suppose, in effect, that event A is repeated
n successive times, while B takes place only once, after the occur-rence of the
last event A. It is clear that if we go by the above defmition, we will have no
reason to consider B as the effect of Ai rather than of Aj ('i' and oj' designating
distinct positive whole numbers not greater than n). We could, of course, require
that the temporal distance between the cause and the effect be determined in a
unique way for each causal relation, since in this case, only one event of the
series of supposedly non-simultaneous events A could be considered as the cause
of B - the one whose temporal distance from B was characteristic of the causal
relation between A and B. We would have to characterize this causal relation in
the following manner: event A, followed after a time t by event B, is said to be
the cause of B if each event A ' having the same intrinsic properties as A, is
followed after an identical lapse of time t, by an event B', intrinsically identical
to B. However, let us suppose that the two events Ai and Aj occur simultaneously
t seconds before B. It is clear that there is nothing to permit us to decide if B
should be considered as the effect of Ai or of Ai> unless we take their spatial
relation into consideration. This example shows that the spatial factor in physical
causality plays a role analogous to that of the temporal factor. If it seems possible
to eliminate the spatial factor from the defmition of the causal relation, the same
must be true of the temporal factor.

This is what is brought out by the principle of action by contact. According to


this principle, event A can act upon event B, separated from it by a fmite
PHYSICAL TIME 199

spatio-temporal distance, only through the intermediary of a chain of events in


which two consecutive envents are contiguous in space and time, the first and
last links of the chain coinciding with A and H respectively. This suggests that it
is preferable to take as our starting point a defmition of immediate cause (which,
in the above definition of the causal relation we will obtain by replacing the
condition governing the existence of a fixed temporal distance between the cause
and the effect with a condition requiring their partial coincidence, i.e., requiring
the existence of an event which is part of both the cause and the effect), and then
derive any causal relation from a fmite number of relations of immediate
causality. We are thus led to the following defmitions:

Event A acts immediately on event H if to each event A I intrinsically identical


to A there corresponds an event H' having the same intrinsic proper-ties as event
H and partially coinciding withA' (3).
Event A acts upon the external 123 event H if it can be linked to H by a chain
of events in which each event acts immediately upon the next event, the two ends
of the chain coinciding with A and H respectively (4).
This last statement can be simplified. Let AI, A 2, An be the chain of events
linking A and H (A 1 = A, An = H). We can see that if A 1 acts immedately upon A
2, and A 2 in turn acts immediately upon A 3, event A 1 will act im-mediately also
upon the event formed by the union of events A 2 and A 3 , which we will
designate by 'A 2 + A 3 '. The same is true of the other events
in the chain. By designating with 'C' the sum of all the events of the chain except
the first (C = ~~Aj) we will be able to replace condition (4) with the
following statement:
Event A acts upon the external event H if there is an event C which con-tains
H and is immediately acted upon by A (5).
The statements (3) and (5) can be replaced with the following single
proposition:
'Event A is the cause of the external event H if there is an event C
which contains H, is reproducible, and partially coincides with
A.' (6).
[In Part IV, below, an alternative defmition of deterministic causality is
developed. H. M.] Thus, the shaking of the windows is considered to be the
effect of a nearby explosion because it is part of the disturbance of the
atmosphere, partially coinciding with the detonation and reproducible along with
it.
This statement requires still another restriction. It is clear that if one adds
200 DURATION AND CAUSALITY

to event A an arbitrary event X, according to (6) the enlarged event A + X will


also have to be considered as a cause of event B, even though it contains events
having no precise connection with B. Now, we would certainly say that
A + X determines B, but not that it is the cause of B, since a cause must
contain only parts which in a certain sense concur in the production of the
effect. But how can we make this sense precise? We cannot require that each part
of the cause be a necessary condition of the effect, i.e. that, if we re-moved
some part of it, the cause should no longer determine the effect and that it should
coincide with a minimal-event which determines the effect, since it is easily seen
that, in general, these minimal-events do not exist. 124 We will require therefore
that each part of A be at least a determining condition
of B, Le., that for every part X of A there should exist an event Y which ceases to
determine B if event X is removed (7). With this reservation we will call A a
cause of B, and we will agree to say that each event contained in the cause of a
given event is causally related to that event. From the reversibility of the causal
laws of physics we have the result that this relation is sym-metrical, Le., that if
event X is causally related to event Y, then Y is causally related to' X. This
amounts to saying that if event X is part of a more extended event X', whose
reproduction entails that of Y, then Y is also part of a more extended event y',
whose reproduction entails that of X.
The symmetry appears here as an empirical property of the causal relation,
resulting from the special form assumed by the causal laws of physics. It would
be otherwise if we defined the causal relation in terms of asymmetrical
causality (implying that the cause is earlier than the effect). We would then have
to say that events A and B are causally related if either one is the cause of the
other; clearly, the relation thus defmed is symmetrical by definition. Yet, in
defming the causal relation in terms of asymmetrical causality we would be
making it depend indirectly upon succession in time, while in statement (6) the
notion of succession has been eliminated. We see that what is essential to the
causal relation is not the fact that one event regularly succeeds another in time,
but rather that the one is reproducible along with the other. Experience
undoubtedly shows that if event A is reproducible along with B, these events are
not simultaneous, but the notion of reproduci-bility is itself independent ofthat of
succession in time.
It will be noticed that defmition (6) of symmetrical causality rests on the
possibility of comparing events from the point of view of the part-whole relation
and from that of intrinsic equality, Le. given two events A andB we should be
able to decide (1) if one ofthem is contained in the other, and (2) if it has the
same intrinsic properties as the other (for example, if they are
PHYSICAL TIME 201

characterized by the same distribution of masses or by the same distribution


of electromagnetic field intensities). The part-whole relation, upon which
rests the first comparison, is a two-term relation which has been studied
axiomatically by Lesniewski 125 ; disregarding the differences in meaning of
the term 'event' discussed in note 120 it is the fundamental relation 'K'
introduced by Whitehead into the study of the spatio-temporal problem. The
equality of intrinsic properties, upon which the second comparison rests also
appears as a two-term relation. But it is clear that if we want to take into
account the continuity of the causal relation, it must be possible to decide not
only if a given event X has the same intrinsic properties as another event Y,
but also whether the intrinsic properties of X more or less differ from those of Y;
it is therefore necessary to be able to evaluate the differences between intrinsic
properties, which, in general, implies the comparison of four events with respect
to one given intrinsic property, i.e., a five-term relation. But, in order to bring out
the logical structure of the causal relation as defmed by statements (2), (6), and
(7), it will suffice to consider the 'degenerate' four-term relation obtained by
supposing that the two intervals under comparison have one term in common; we
will say that event X, which differs from events Y and Z with respect to intrinsic
property c/J less than they differ from each other in this respect, is between
Yand Z with regard to this
property, and we will designate this four-term relation by the latter 'F. The
expression 'I(X, Y, Z, c/J)' is read 'X occurs between Y and Z with respect to
c/J', which means that event X differs from event Y and Z less than they differ
from each other with respect to the intrinsic property c/J.
The causal relation defmed above can be expressed in terms of the relations
I and K. In order to show this we will put forth the following defmitions: We
shall first defme the intrinsic properties of events in terms of their capacity
to become the fourth term in the relation I. More precisely: c/J is an intrinsic
property if there are at least three events X, Y, and Z such that I(X, Y, Z, c/J)
is the case (8). The two events, X and X', exhibit the same distribution of the
intrinsic property c/J if the propositions I(X, Y, Z, c/J) and I(X', Y, Z, c/J) are
equivalent for every pair of events Y and Z (9). Since this equality of distri-
butions is by definition a symmetrical and transitive relation, we can defme
'by abstraction' the distribution of an intrinsic property as the set of events
having the same distribution as a given event (10). The different distributions
of a single intrinsic property are arranged in a qualitative order derived from
that of the events: we will say that the distribution ~ of the property c/J lies
'between' the distributions 1/ and t of c/J if three events X, Y, and Z, which
respectively exhibit these distributions, fulfill the condition I(X, Y, Z, c/J)
202 DURATION AND CAUSALITY

(11). Let E and E' be two sets of distributions of the two properties cp and
q/, which can also coincide: we will say that an intrinsic function F is defined
in the set E, if, to each distribution contained in E, there corresponds a single
distribution contained in E'; symbolically, we write this '~' = F(~)' (12). In
order to delimit the continuous functions we will use the notion of conver-
gence: we will say that the distribution ~ is the limit of the convergent series
of distributions ~ 1, ~2' ... ~n, ... of an intrinsic property cp (symbolically, 'lim
~n = r) if each term of this series occurs between each preceding term and ~ (in
other words, if one hasI(~m+n' ~n' ~,cp,) for each whole positive m and n) and if
there is no distribution lying between all the terms of the series and ~ (13). A
function ~' = F(~) is continuous if the equality lim ~n = ~ always entails the
equality lim F(~n) = F(~) (14). In the same way we shall defme the continuity of a
function of several distribution variables, each concerning a different intrinsic
property. In using definitions (8)-(14) we will replace statements (2) and (6) with
the following defmition: event A is the cause of event B if there exists a
continuous intrinsic function ~ = F(~, 'T1 ... ) such that an event C, partially
coinciding with A and containing B, fulfills the equality k = F(~A' 'T1A' ... ), and
that each event A' partially coincides with an event C' satisfying the equality ~C'
= F(~A" 'T1A" ... ) (15).

The causal relation acquires simple properties only in its application to


elementary events, which last an. infmitesimallength of time and occupy an
infmitesimal amount of space.126 The different causes of a given elementary
event are separated from it by a variable temporal distance. An important
property of physical causality is that the set of determining conditions of an
elementary event, realized at a given instant, always constitute a cause of this
event: the set of determining conditions of a present event, having been
realized an hour previously, make up a complete cause of that event, just as
do the sets of conditions realized a second or a century before. Each
elementary event therefore has an infmte number of causes in its past, but
only one cause at any determinate time in its past. Thus, the present position
of a speck of dust, which we can take as approximating an elementary event,
was predictable a second, or a minute,or an hour ago. Someone who,a second
ago, had wished to predict the position of the speck of dust would have had
to know all the contemporaneous events capable of exerting an influence
upon its future position. He would have had to ascertain, for example, that
there was no body within one meter of the speck which was moving toward it
with a velocity greater than one meter per second, since the presence of such
a body would have been capable of exerting a decisive influence on the
PHYSICAL TIME 203

position of the speck; more generally, he would have had to ascertain that
there was no body within x meters of the speck which was moving toward it
with a velocity greater than x meters per second. Must we conclude that,
since x is an arbitrary number, it would have been necessary to know at that
time the distribution of matter throughout all of space in order to predict the
position of the speck of dust one second later? From the point of view of the
theory of absolute time, according to which any event earlier than a given
event is its determining condition, this conclusion is well founded. But it is
not so from the point of view of relativity, which implies the existence of a
limiting velocity for matter, and from which it follows that in order to predict
the position of the speck at the end of one second, it is sufficient to know the
distribution of matter in the region of space from which a body could reach
the speck without exceeding the limiting velocity, i.e., the distribution of
matter in the sphere whose center is the present position of the speck of dust
and whose radius is equal to the limiting velocity of matter. If we wished to
predict the position of the speck in x seconds, we would have to know the
distribution of matter in a sphere with a radius x times as large. We see that
the cause of an elementary event, located at a determinate moment in the
past, occupies a greater region of space the more distant the moment from
the date of the event. However, this region never fills all of space. From the
point of view of the causal theory of time, this is an extremely important
consequence of relativity, since it considerably simplifies the problem of
predicting the future.
In view of the enormous value of the limiting velocity, however, this
simplification is only theoretical. The prediction of an event one second away
would require the know ledge of the state of a sphere with a radius of 300,000
kilometers, which will always be unrealizable in practice. In practice, this
difficulty is resolved in two ways, which are essentially equivalent: we re-nounce
either the total prediction of the future, or the certainty of the prediction. In the
first case we try to know the initial conditions in a limited domain accessible to
observation, and then try to estimate the conditions at the boundaries of this
domain during the relevant time interval separating the instant of the prediction
and the instant at which the event in question will take place; thus, in order to
predict the future state of some mechanism, it sufficies to study its present state
and estimate what will take place at the surface of the region enclosing the
mechanism during the time interval sepa-rating the future state from the present
state. In the second case, we hypothe-tically complete the knowledge of the
initial data by making the general supposition that the initial conditions external
to the restricted domain
204 DURATION AND CAUSALITY

accessible to observation do not contain events capable of exerting a signi-


ficant influence upon the coming event. We see that the two procedures do
not differ essentially from each other: estimating the boundary conditions is
equivalent to completing the initial conditions in such a way that the latter
engender the former.
We conclude from these considerations that defmition (15) of the causal
relation, stated on page 202, proves that the causal relation is theoretically
defmable without reCOlfse to the notion of time, but not that it is experi-
mentally possible to ascertain a causal relation without utilising perceptions
of the temporal order. Notice, however, that often perceptions of the temporal
order linking events of moderate extension in the neighborhood of the
observer and involving only velocities negligible with respect to him, suffice
to establish with a practically satisfactory degree of probability a causal
relation between events which are distant from one another or which involve
velocities and extensions inaccessible to direct observation, and thereby
deter-mine their temporal orientation. This fact is important in the
epistemological interpretation of the causal theory of time, since it brings out
that it is not the aim of this theory to replace the perceptions of the temporal
order every-where with a study of causal interconnection, but only to
generalize these perceptions by extending the local temporal order, revealed
in perception, to regions inaccessible to it, or merely unperceived, and thus
inserting it in cosmic time.
Thus, the departure of a radiotelegraphic or other signal from a given region
and its arrival in another region form two events whose spatio-temporal structure
can be determined by local observations. These observations bring out the
intrinsic resemblance between the two events and thereby make their
independence highly improbable, but, taken by themselves, they do not
guarantee that it is a matter of the departure and arrival of the same signal. Here
we hazard certain hypotheses about the boundary conditions accom-panying the
trajectory of the first signal, supposing, for example, that having been launched
in an often explored direction, it should not have encountered obstacles which
would have turned it from its original direction toward the second region. These
are the hypotheses, some of which we occasionally verify, which lead us to the
belief in a causal connection between the two locally observed signals, and
which allow us to determine their temporal orientation, which results from this
connection. Undoubtedly, the knowledge of this interconnection always remains
hypothetical, as we can see from the preceding remarks on the prediction of
physical events. But the knowledge of the temporal order of physical events, as
distinct from the order of their
PHYSICAL TIME 205

perceptions, is also hypothetical, and it is with the aid of the hypotheses on the
causal interconnection of phenomena that we determine their temporal order.

3. THE CONCEPT OF EVENT 127

We shall now consider the terms that the causal relation is said to lin1e In our
axiom system, these terms are called 'events', which amounts to defming the
event as· an object linked to other objects by the causal relation. This is
obviously only a linguistic matter, but the important thing is to know to which
otherwise known concept this notion of event corresponds. The axioms attribute
to each 'event' the property of occupying only an infmitesimal amount of space,
an infmitesimal amount of time, and the capacity of coin-ciding in time and
space with other events. This shows that the event of the system corresponds not
to the world-point, since two world-points which coincide are identical, but to
the elementary event. Thus, the instantaneous state of a material point can
coincide in space-time with the instantaneous state of a point of the
electromagnetic field: these two elementary events occur at the same world
point. A world point may be defmed as a set of elementary events coinciding in
space-time with a given elementary event.
The affirmation contained in the system - that the causal relation is supposed
to link elementary events - may suggest several objections which it would be
useful to consider at the present time. These objections relate to the very notion
of elementary event, and also to the idea that the causal relation links events. Is it
not contradicted by Schopenhauer's well-known opinion that causality links only
changes (Le., certain special processes), and also by Kant's opinion (contested by
Schopenhauer) that the causal relation occurs between things, since it is things
which act on one another?
We shall not be concerned with the question of whether, from the point of
view of the universe of common sense, causality should be considered as linking
things, processes, or events. Since, from this point of view, the three notions seem
to be mutually irreducible, as Ingarden has shown, 128 it is quite possible that,
applied to the universe of common sense, this ·question has a determinate sense.
The same is not true if we apply it to the physical universe, where 'things' and
'processes' are defmable as a function of events. The existence of causal relations
among physical events involves the existence of corresponding relations among
the processes and things formed of these events, and moreover, these relations
among processes and things have properties different from those of the causality
of events. The question of
206 DURATION AND CAUSALITY

knowing which of these relations should be considered as most fundamental with


respect to the others does not make much sense. In our axiomatic theory, it is the
causal relation among events which occupies this position, but this evidently
depends upon the choice of primitives.
But, basically, what does the term 'event' mean, in its physical sense? It is
generally taken as a simple indefinable expression. Yet, subsequent remarks seem
to show that it does not have the same meaning for different authors. For us,
'event' is not opposed to 'thing'. The durable thing is a set of un-extended and
instantaneous things: the unextended and instantaneous thing is an event. We
therefore see no opposition between the point of view of 'substantival subjects'
and that of 'events', discussed by Zawirski 129 in connection with Whitehead's
theories. It is true that Whitehead opposes 'event' to 'thing'. Yet his examples of
events are examples of things: "The event which is the passage of the car is part
of the whole life of the street. Also the passage of a wheel is part of the event
which is the passage of the car. Similarly, the event which is the continued
existence of the house extends over the event which is the continued existence of
a brick of the house and the existence of the house during one day extends over
the existence during one specified second of that day." 130 The objects mentioned
in these sen-tences are essentially nothing but material things: street, car, wheel,
house, brick. What gives the impression that they are radically different objects is
only the substitution of abstract substantives for the verbs: instead of a passing
car, it is the passage of a car, instead of an existing house, it is the existence of
the house. It is difficult to believe that such a banal linguistic transformation
could bring into being a whole new and irreducible reality of events alongside the
reality of things, (and in fact, Whitehead makes no such claim).

Does this mean that it makes no difference whether we speak of things or of


events? Certainly not, since there are elementary events which are not
instantaneous things, if it is understood that every instantaneous thing can be
prolonged into an enduring thing. The proposition: 'at a certain instant, at a
certain distance from a certain body, there was a certain intensity of the
electromagnetic field' certainly refers to an individual event, even though this
event does not necessarily coincide with any enduring thing: the electro-magnetic
field is also dermed in the 'void'. Moreover, this proposition is susceptible of
several interpretations: (1) The intensity of the field at instant T at point P is
considered as the state of a particle of the ether which coin-cides at this moment
with this point in space. This is Faraday's primitive interpretation implying the
notion of a substantival ether, which had to be
PHYSICAL TIME 207

abandoned when Relativity appeared. (2) The intensity of the field is a property
of the spatial point and varies with time. This primitive interpreta-tion, although
not utilizing the notion of the ether, does depend upon the hypothesis of absolute
space, refuted by Relativity. (3) The intensity of the field is a property of the
world point, conceived as an element of space-time. This interpretation assumes a
substantialistic theory of space-time, according to which the world point would
be an ultimate reality. It is rejected by the relational theory, and consequently by
the causal theory of time, since the latter is only a special case of the former. (4)
The intensity of the field is a property of an unextended element of the field, of
an elementary event. The field is considered here as a set of real elements which
are neither instants of time nor points of space; space and time are considered as
sets of well-defined relations among all kinds of events, those constituting the
history of the electromagnetic field and those constituting the history of matter.
Since it permits us to formulate the axiomatic theory of space-time in the
simplest possible way, we will accept this fourth interpretation. The instantaneous
and unextended events constituting the history of matter and the electromagnetic
field are considered here as fundamental entities, of which only the first are
arranged in a series capable of being taken as the history of a single particle
retaining its identity through time. [In Part III below, we will introduce quan-tum
theoretical evidence to show the need for a language whose basic universe of
discourse consists of space-time regions while events are assigned a higher
logical level. H. M.]

There still remain a few words to be said on the infinitesimal character of


elementary events. It is clear that such events are not directly accessible to
observation, which always yields data of finite spatio-temporal dimensions. This
is why Whitehead undertook to reduce all physical notions which imply
elementary events to notions implying only fmite events. The ingenious 'method
of extensive abstraction', which he developed for this purpose, in fact permits
making each statement concerning an unextended event corre-spond to an
equivalent statement concerning an infinite series of events extended in space
and time. This method is applicable in principle to our axiomatic theory, and
could be used to replace the causal relation among elementary events with the
causal relation among fmite events. For this purpose it would suffice to take as
primitive the causal relation among extended events and then define the part-
whole relation in terms of it, say-ing, for example, that one event is a part of
another if each effect of the former is also an effect of the latter, but not vice
versa. The concept of an elementary event could then be defined within the
system by the method
208 DURATION AND CAUSALITY

of extensive abstraction. It is true that, alongside the causal theory of time,


this method would seem repetitious, since Whitehead claims, with the aid of
his method, to be able to deduce not only the definitions of unextended
events in terms of extended events, but also the topology of space-time, and
therefore, in particular, the characteristic properties of the temporal order.
Yet, he succeeded only with the aid of arbitrary postulates, such as the one
implying that the only infmite events are the 'durations', i.e., the segments of
becoming which include all the elementary events contained between two
states of the universe. Moreover, in the course of his work, Whitehead is
forced to introduce a second primitive notion, that of 'con-gredience', which
includes both the notions of simultaneity and of spatial coincidence.

Therefore, I believe that the only advantage of the method of extensive


abstraction is to link the notion of elementary event to that of finite event.
But this result can be reached by much simpler means. What is essential in
the epistemological interpretation of physics is to represent the physicist as
operating on entities which correspond as directly as possible to empirical
data. To us there seems to be no doubt that he does not operate directly on
these data. The elementary event does not appear to meet this condition,
since sensory data are always extended. For example, when I see a box of
matches, the sensory datum is a colored surface which is also endowed with
a certain temporal thickness; the physical entity which is supposed to
correspond to this sensory datum is a spatio-temporal region exhibiting a
determinate distribution of masses and field vectors. This region is charac-
terized not by a single elementary event but by the set of all the elementary
events contained in it. Therefore what corresponds to the sensory datum is
not a fundamental entity, but a suitably delimited set of these fundamental
entities of our axiomatic theory. We see no drawback in this. Undoubtedly,
the correspondence would be more direct if it were the spatio-temporal
region which figured as the fundamental entity of the system. However, this
small complication is counterbalanced by the enormous simplification in the
formulation of physical laws which results from the use of elementary
events, and consequently it seems to be completely justified.
Moreover, it is clear that the part-whole relation linking spatio-temporal
regions conceived as sets of elementary events has properties analogous to
those of the homonymous relation existing between extended events. It is
essentially the same relation represented differently in two different axio-
matizations of physical science. We have already pointed out 131 that the
different axiomatizations of a given set of propositions involve variations
PHYSICAL TIME 209

not only in the groups of supposedly undemonstrable propositions and in


supposedly primitive notions, but also in the logical types of the variables
employed. This is true of the part-whole relation in both Whitehead's axio-
matization and our own.

4. THE ORDINAL CONCEPTS

4.1. Coincidence

Two events are said to coincide in the space of a frame of reference if, from
the point of view of this frame, they occur at the same place; they coincide in
the time of the frame if, from its point of view, they occur at the same
instant. Events which coincide in the space and time of a given frame of
reference also coincide in the space and time of every other frame. Spatio-
temporal coincidence is therefore an invariant relation, while spatial coin-
cidence and temporal coincidence of two events in general take place only
relative to a determinate frame of reference. Obviously, it is possible to
defme spatio-temporal coincidence as a conjunction of spatial coincidence
and temporal coincidence,132 but it seems preferable here to give this
important concept an independent defmition.
Two coincident events are considered as taking place at the same world
point, or occupying the same position in time and space. But, according to
the very principle of the causal theory, the causal relations linking a given
event to the set of other events determine what position in space-time is
occupied by this event. Because this position is identical for the two events
in question, the relations of causality which link them to the set of other
events must also be identical. We are thus led to the following definition:
Two events coincide in time and space if they have all their
effects in common. (D 1)
The spatio-temporal coincidence of a pair of events is a very close relation,
but it does not involve their numerical identity. A world point can therefore
contain several events. However, this difference between the event and the world
point will not become important until section 11; until that time it will be more
convenient not to distinguish spatio-temporal coincidence from numerical
identity, i.e., to identify the event with the world point.
The fundamental properties of coincidence follow immediately from its
defmition: (a) Each event coincides with itself. (b) If event A coincides with
210 DURATION AND CAUSALITY

event B, event B coincides with event A . (c) Two events which coincide with a
third event also coincide with each other.
The notion of spatio-temporal coincidence has been particularly favored by
many of the interpreters of relativity. Some, following a remark of Min-
kowski,133 consider every physical phenomenon as a coincidence of world lines
and reduce the spatio-temporal structure of becoming to an 'order of coincidence'.
Others insist upon the epistemological role of coincidence; according to them
observations bear only on coincidences (such as the per-ception of the
coincidence of the needle with a number on the face of the clock). The first
interpretation, which is of an ontological nature, does not appear to be very clear.
If it means that every physical phenomenon is a coincidence of world lines of
material particles, one might object that these lines never coincide. If it is a
matter of arbitrary world lines, an infinity of them could be formed passing
through any event, and it is not at all clear why the event must be considered as a
coincidence of all these lines instead of regarding the lines as sets of events. From
the epistemological point of view, the alleged primacy of coincidence is also
questionable. Undoubtedly, the coincidence of two sensory objects (e.g., that of
the needle with a number on the face of the clock) often seems to be an
immediate and indubitable datum; but so are the spatial arrangement of the
numerals on the face, the before and after in their movement, the inside and
outside of their parts. Generally, any local spatio-temporal order is perceivable in
principle, and coincidence has no privileged place; in the universal spatio-
temporal order, which results from an indirect generalization of this local order,
coincidence plays a part similar to that of the other ordinal concepts.

4.2. The Spatia-Temporal Order


The spatio-temporal order expresses those properties of space-time which are
independent of all measurement. We can define it in terms of the causal relation
by utilizing the principle of action by contact. This principle implies that if two
distinct events A and B have an effect in common, this effect must have traversed
the spatio-temporal domain separating them. Besides, this domain will be filled
with the effects of the events taking place in the interval separating A and B.
Thus, the events situated between A and B will influence all the effects common
to A and B. These remarks suggest the following definition of the temporal
order:
Event B takes place between events A and C if each effect
com-mon to A and C is also an effect ofB.
PHYSICAL TIME 211

We shall also say in thls case that B occurs in the interval separating A
and C, that C occurs outside A and B, and that A occurs outside Band C 134 (D
2).
Notice that the order thus defmed is in itself neither temporal nor spatial.
It will become spatial in the frame of reference in which A and C (and con-
sequently also B) are simultaneous; it will be temporal in a frame of reference in
whlch A and C succeed each other in time. The fact that event B takes place in
the interval separating events A and C has an invariant significance, whlch may
be expressed grosso modo by saying that events which are closer together have
more effects in common - which once more brings out the kinship between the
causal definition of the spatio-temporal order and the principle of action by
degrees.
Notice further the isotropic character of the order just defmed: it follows
directly from this defmition that an event which occurs between A and B
occurs also between B and A; no absolute orientation of the pair A, B is
therefore assumed. If we take into account the definition of coincidence, we
see that the spatio-temporal order of events is unchanged by coincidence and
therefore depends only on the position occupied by the events in space-time:
every event which coincides with an event taking place between A and B
also takes place between these events.

4.3. Simultaneity
A causal decomposition is a decomposition of all events into mutually
disjoint classes satisfying the following conditions:
(1) events of the same class are causally unrelated,
(2) an event not belonging to a given class has at least one effect in
that class,
(3) an event taking place between or outside of two events of a given
class also belongs to that class.
These classes are called instants of the decomposition. Two events be-
longing to the same instant of a decomposition are simultaneous in that
decomposition. (D 3)
This defmition can be rigorously justified (cf. Chapter II, Section 10).
Condition (1), whlch expresses the non-existence of the causal relation among
simultaneous events, or, what amounts to the same thing, the non-existence of an
infinite velocity of propagation, will be discussed on pp. 215-16. Con-dition (2)
asserts that every event possesses determining conditions at each
212 DURATION AND CAUSALITY

epoch of its past and future. Condition (3) is less familiar - it could be
compared to the property which the plane has of containing any straight line
which passes through two of its points. The classes which constitute a causal
decomposition are not planes, but rather instantaneous spaces. Condition (3)
therefore amounts to asserting that an instantaneous space contains every
straight line passing through two of its points. It goes without saying that
these remarks, peripheral to the system, must add nothing to the definition of
simultaneity, which is derived from the concept of the causal relation.
The transitivity and symmetry of simultaneity follow immediately from
definition D 3. The relativity of simultaneity does not: definition D 3 would
be equally valid in the theory of absolute time, provided that this theory
admitted the non-existence of an infinite velocity of causal propagation. 135
But neither does this definition exclude the possibility that two events
occurring at the same instant of a certain decomposition might not occur at
two different instants of another decomposition. Notice, however, that two
coincident events are simultaneous in every decomposition. In fact, if events
E and E', assumed to be coincident, belonged to instants t and t' respectively
of a given decomposition, according to (2), event E would have at least one
effect at instant t'. This effect would also be an effect of E' (since, by
defmition, E and E' have all their effects in common); there would then be at
instant t' two events linked by the causal relation, contradicting
(1).
Definition D 3 raises some methodological problems. It is now common-
place to assert that only definitions which provide experimental procedures of
verification are accepted by the physicist and have some physical meaning. Now,
definition D 3 makes the simultaneity of two events depend on the behavior of
the whole universe. To decide if events A and B are simultaneous it would be
necessary first to form a causal decomposition of all the events and then check to
see whether these two events happen to be in the same instant of this
decomposition. Thus formulated, the objection is not correct, since the question
of whether two events chosen at random are simultaneous is meaningless one -
the same pair of events can be simultaneous or not, depending on what
decomposition you choose. What is important is to find out if two given events
are simultaneous in a given decomposition. But what is a given decomposition?
If, in order to determine it, it were necessary to know at what instant each event
took place, it would obviously be impossible to establish any simultaneities. But
one result of the axioms which follow is that a causal decomposition is
determined, in a general way, by four events. 136 (More precisely: if four events,
not in the same place, are simultaneous in
PHYSICAL TIME 213

two decompositions, these two form only one). The question, therefore, is to
know whether events A and B are simultaneous in the decomposition in
which four other given events, C, D, E, and F, are simultaneous. This
evidently involves only one relation among six determinate events, and not
one between A and B and the entire universe.
The definition of simultaneity must provide a necessary and sufficient
condition for this relation, and not just a convenient criterion. Its applic-ability
will be assured if there are cases where this relation of six events can be
ascertained on the basis of a finite number of data, without bringing in the whole
universe. The existence of such cases follows from the fact that we can, by using
the axioms of the system, deduce from definition D 3 some appropriate criteria of
simultaneity, for example, Einstein's optical criterion.137 In those cases where
the events in question are linked by light messages, their simultaneity can be
established without introducing the set of all events. Furthermore, even in certain
cases where optical signals (or, more generally, electromagnetic and material
ones) are lacking, it is possible, as the examples given above suggest, to draw a
criterion of simultaneity from definition D 3 by considering a decomposition of a
limited domain containing the six events in question instead of a decomposition
of all the events.

4.4. Succession
Two events which are not simultaneous in a given decomposition are said to
succeed each other in the time of that recognition. Thus conceived, succession is
a symmetrical relation, since it is equivalent to non-simultaneity. It is therefore
not adequate for the definition of a temporal order, which, as Leibniz said, is
precisely the order of successive events. Ordinarily, it is the asymmetrical
relation of before and after which is used to define the temporal order. However,
we have seen that the distinction of before and after seems devoid of any intrinsic
physical meaning,138 and that by basing the axiomatic theory of physical time
on asymmetrical succession we would, by the same token, suppose the question
of reversibility to be answered in the negative. It is not intuitive propositions of
the type, 'A is before B', but rather those of the type 'B occurs in the time
interval separating A and C' which seem susceptible of a causal generalization.
We accept the following definition:
Event B takes place in the time interval separating events A and C,
relative to a given decomposition, if there are three events A', B', and C',
respectively simultaneous with events A, B, and C in this decomposition, and
such that B' takes place between A ' and C'. (D 4).
214 DURATION AND CAUSALITY

The order of the events A', B', C' is that of definition D 2. We conclude from
this that the order of the succession is also isotropic: an event taking place in the
time interval separating A and C also takes place in the time interval separating C
and A. The two ends of the temporal interval play per-fectly symmetrical roles
here.
By basing the succession of events upon that of instants, we can put
definition D 4 into an equivalent but more intuitive form. We will say that instant
t2 lies in the time interval separating instants t 1 and t 3 if it contains an event B
taking place between two events, A and C, which are located at instants t1 and t3
respectively. The temporal order of events will then be identical to the order of
the instants to which they belong.
It is clear that these equivalent definitions define a relative order: the events
which succeed one another in a certain decomposition will be simul-taneous in
some other, on condition, however, that there be no causalrelation among them,
since, in this case, according to D 3, they cannot be simul-taneous in any
decomposition.

5. THE FIRST GROUP OF AXIOMS


I 1. There are at least two events which are not causally related.
I 2. To each pair of causally unrelated events there corresponds at
least one decomposition in which they are simultaneous.
I 3. To each pair of causally unrelated events A, A', taking place at
instants t and t' respectively of a given decomposition, there
corresponds at every other instant of this decomposition at least
one event which takes place either between or outside of A and A'.

Axioms I 1-2 guarantee the existence of one decomposition; Axiom I 3 is the


isotropic counterpart of the property of asymmetrical succession, con-sisting in
the fact that of any two instants of a decomposition, one is always before or after
the other. These axioms imply the following theorems:

THEOREM 1. For two events to be causally related, it is necessary and


sufficient that there be no decomposition in which they are simultaneous
(I 2, D 3).

THEOREM 2. No event is its own effect.


Proof. According to I 1-2, there is at least one decomposition. In this
decomposition every event is simultaneous with itself. Therefore, if there
PHYSICAL TIME 215

were an event which was its own effect, this event would be simultaneous with
its effect, contrary to D 3.
THEOREM 3. If event A is an effect of event B, event B is also an effect of
event A.
Proof. If B were not causally related to A , there would be a decomposition
in which these two events were simultaneous, although one of them was an effect
of the other, contrary to D 3.
It follows from Theorems 2-3 that the causal relation is not transitive. This
becomes obvious if we replace the expression 'causal relation' with the
alternative one 'cause or effect': from the fact that event A is a cause or an effect
of event B, and that event B in its turn is a cause or an effect of event C, we can
conclude nothing about the existence of a causal relation between events A and
C. This fact is characteristic of our system of axioms. By making use of an
asymmetrical and transitive relation, Carnap, Reichenbach, and Robb thereby
accepted an anisotropic temporal order. This is not so if we take a symmetrical
and intransitive relation as our point of departure, since only an isotropic order is
then definable in terms of it. To introduce asym-metrical succession a new axiom
would be necessary. eCf. pp. 235-236J
THEOREM 4. Two events which coincide are never causally related.
Proof. If event A and B coincide, each effect of one is also an effect of the
other. Therefore, if B were an effect of A, it would follow that it would also be an
effect of B, contrary to Theorem 2.
This consequence may invite criticism; one readily accepts the non-exist-ence
of causal relations between distant simultaneous events, since their existence
would imply an infinite velocity of causal propagation. But a connection between
coinciding simultaneous events would not go against the principle of action by
degrees. Most of the fundamental laws of physics are precisely about a functional
dependence among magnitudes corresponding to the same worked point. The
Maxwell-Lorentz equations link the local values of the electromagnetic field
vectors to the density and local velocities of electric charges; Einstein's law
relates the local value of the energy-momentum tensor to the local values of the
metrical tensor and of the curvature. Simi-larly, in the statistical realm, the law of
ideal gases seems to link the local values of the three magnitudes which
characterize the gaseous state. But the question is precisely whether it is
convenient to call these functional relations causal relations. If this were the case,
it would be necessary, in the definition of a causal decomposition, to replace the
condition requiring the non-existence
216 DURATION AND CAUSALITY

of any causal relation among those events which are members of the same
instant, with a more restricted postulate requiring only the non-existence of
this relation among the non-coincident events of the instant. Although it
would eliminate the theorem concerning the non-existence of any causal
relation among coincident events, this change would not greatly alter the
structure of the system. The language of physics is undoubtedly hesitant on
this point, and only a convention could resolve the issue. Nevertheless, we do
believe that we are in better accord with the tendencies of this language when
we refuse to consider these functional relations among simultaneousmagni-
tudes as causal relations. It seems to us that no physicist would hesitate to
affIrm that, for there to be a causal relation between two events, it should be
possible, at least theoretically (if not technically), for the events to be linked
by a signal. But it is clear that two events which coincide in space and time
cannot, even theoretically, be linked by any signal. This theoretical possibility
is undoubtedly very vague, but it nonetheless reflects a certain tendency of
the language of physics.

6. THE GEOMETRICAL CONCEPTS

The time of a decomposition is the set of its instants; the space of a decom-
position is the set of its points. Just as the instant was defined, without
circularity, as the class of events which occur at the same time as a given
event, so the spatial point may be defined as the class of events occurring at
the same place as a given event. The property of events to occur at the same
place is a relative property, as is that of occurring at the same time. Further-
more, this relativity of spatial coincidence was clearly known before
Einstein, and is far from being as paradoxical as that of temporal
coincidence. Thus, two successive coincidences of the hand of a clock at rest
with respect to the Earth with a numeral on its face, occur at the same place,
from the point of view of the Earth, but, viewed from the Sun, they are
separated in space by a distance of several kilometers. We accept the
following definition of spatial coincidence :
Events A and B occur at the same place in a given decomposition if, in
this decomposition, there exists an instant t, containing neither A nor B, and
such that every effect of A at t is also an effect of B, and vice versa (D 5).
We shall see later that two events, which are spatially coincident but non-
simultaneous in a given decomposition, are always causally related. But the
converse is not true. There are events which are causally related but which are
not spatially coincident in any decomposition. We shall call each effect of an
PHYSICAL TIME 217

event which does not coincide with it in the space of any decomposition a
limiting effect of this event. It is clear that if event A is a limiting effect of event
B, B is also a limiting effect of A. The pairs of events of which one member is a
limiting effect of the other are of capital importance for the structure of space-
time (pairs of events belonging to the history of a single light ray are always of
this kind). They allow us to define spatial equality.
Let A, B, C, D, be four simultaneous events in a decomposition 8. We will say
that in this decomposition the distance from A to B is equal to that from C to
D if a limiting effect of A coinciding with B in the space of 8 is con-
temporaneous with a limiting effect of C coinciding spatially with D. (D 6).
Taking into account the optical signification of limiting effects and of
Theorems 6 and 8, we see that these two definitions lead to the following criteria
for spatial coincidence and congruence : (a) Let A be the departure of rays
emitted by the explosion of a small body located at the center of a spherical
material surface; let B be the meeting of these rays after they have been reflected
by the inner walls of the sphere. Events A and B will coincide by definition in the
space of the decomposition in which the sphere is at rest.
(b) Let A and A I be two explosions occurring at instant t of a decomposition
8, and let Sand S' be the two spherical surfaces lighted at instant t' ofD by the
rays emitted by A and A' respectively. By definition, the radii of these spheres
will be equal in the decomposition 8. It goes without saying that these statements
are only optical criteria of spatial coincidence and congruence, as these two
relations are not restricted to spatio-temporal regions traversed by light.

The concepts of spatio-temporal order and of congruence suffice to defme all


the other fundamental concepts of geometry. The point is represented by the
event. A straight line passing through points A and B, assumed to be
simultaneous in the decomposition in question, by definition will contain,
besides A and B, all points X such that: either A is between X and B, or
X is between A and B, or B is between A and X. We can see that the axioms of
geometry can be considered as expressing the properties of the causal relation.

7. THE SECOND GROUP OF AXIOMS

II 1.At each instant there is only one event occurring at the same place as a
given event.
II 2.If events A, B, C, which occur at instant t of a decomposition 8, occur
respectively at the same place as A', B', C', occuring at
218 DURATION AND CAUSALITY

instant t ' of 0, and if B occurs between A and e, then B' occurs


between A and e'.
I

II 3.139 If events A, B, e, D, occurring at instant tofa decomposition 0,


occur respectively at the same place as eventA', B', e', D', which
occur at instant t ' of 0, the equality of AB = CD entails the
equality A 'B' = eID'.
II 4.In each decomposition there is at least one instant at which the
geometry is Euclidean.

Axioms II 1-3 express the fact that the one-to-one correspondence


between two instants of a decomposition, which is defined by the spatial
coincidence of their elements, leaves the ordinal and metrical relations among
these elements invariant. We conclude from this that geometry, being Eucli-
dean at one instant of a given decomposition, cannot but continue to be so at
every other instant, by virtue of the properties of spatial coincidence. All
these instantaneous geometries define an enduring geometry characterizing
the space of the decomposition. This space is made up of enduring points
each of which is a set of events coinciding spatially with a given event. Order
and congruence in the enduring space are defined by a combination of the
instantaneous order and congruence, as defined by D 2 and D 6 with spatial
coincidence; we shall say, for example, that the enduring segment AB is equal
to the enduring segment CD if there are four simultaneous events, A B', e', I,

D', located at the enduring points, A, B, e, D, respectively and satisfying the


equality A 'B' = eID', in conformity with defmition D 6. The enduring geometry
is, moreover, no less relative than the instantaneous geometries it contains - each
decomposition has its own space. The events which coincide in the space of one
decomposition do not always coincide in the space of another.

It will be noticed that Axioms II 1-3 would be satisfied in the decomposi-tion


in question even if there were only one instant, if its time were reduced to an
instant. But it follows from II 4 that this is not the case. From it we see that if B
occurs between A and e, A does not occur between Band C. It follows that there
are at least two events which are causally related. In fact, the proposition that B
occurs between A and e means that every effect common to A and e is also an
effect of B. If none of these three events had any effects, it would follow that A is
also between Band e, contrary to 114. One of these events, say A, has therefore
at least one effect,A'. Since A and A are not simultaneous in any decomposition,
I

the decomposition in question must contain at least two instants, t and t', which
contain the events A and A I
PHYSICAL TIME 219

respectively. Let us take event B to be contemporaneous with A. I say that


there must be at the instant t' at least one event which is not a common effect
of A and B. In fact, if this were not the case, according to Definition D 5,
events A and B would occur at the same place; therefore, at instant t there
would be two distinct events occurring at the same place as event A, contrary
to II 1. There are therefore two non-simultaneous and causally unrelated
events, say Band B'. According to I 2, these events will be simultaneous in
another decomposition, 0', where, according to II 4, they will determine a
straight line. No two events located on this line are simultaneous in 0, since,
according to the definition of a straight line and to Definition D 3, this
simultaneity would entail the simultaneity of all the other events on the line,
contrary to our hypothesis that Band B' are not simultaneous. On the other
hand, according to I 3, each instant of 0 contains at least one event located on
this line. We therefore obtain a one-to-one correspondence between the set of
instants of the decomposition 0 and the points on a Euclidean straight line of
decomposition 0'. Since these two decompositions were chosen arbi-trarily,
we can state the following theorem:
THEOREM 5. All the instants o/an arbitrary decomposition/orm a linear
continuum. 0/ three instants 0/ a decomposition, there will always be one
which occurs in the time interval separating the other two. We shall see
that this order 0/ instants is independent 0/ the particular choice 0/
events B, B'; cf. THEOREM 11. p. 222.
Axioms I-II describe certain properties of time and space, without es-
tablishing a hierarchy between these concepts: the spatial axioms follow
those of simultaneity and serve in tum to derive the temporal properties of
non-simultaneous events (the order of instants is found to be derived from the
linear spatial order). Shall we say that this procedure does not take into
account the natural hierarchy of these concepts, since almost all the
supporters of the causal theory of time (Leibniz and Kant implicitly, Camap,
Reichenbach, and Robb explicitly) have maintained that the concept of time
is more fundamental than that of space? But on what would this hierarchy be
based? Ontologically, the possibility of deducing the properties of space from
those of time does not prove it, since we can just as well define the latter in
terms of the former. Epistemologically, causal time seems to us to be just as
far removed from immediate experience as is causal space; intuitive time is
an empirical datum on a par with intuitive space.
Nor can we accept Bergson's diametrically opposed opinion, according to
which physical time is merely derived from space, as an additional dimension
220 DURATION AND CAUSALITY

(as he said in his Essai sur fes donnees immediates de fa conscience,


long before Relativity, [this view was already presented by Lagrange. H. M.J)
for the role of time in the order of events is not assimilable to that of space.
Bergson's thesis might suggest a purely epistemological statement, to the
effect, that, since only the present is given to us, we must construct physical
becoming in terms of it, which is basically the same thing as deducing the
temporal order from the spatial order. This physical present-ism seems
unfounded to us. The physical present is constructed out of sensory data, as
are the past and the future. These data (such as the sensations corresponding
to distant objects) are often more directly related to the distant past than to
the present. The only real epistemological priority seems to us to be that of
the causal relation with respect to the time and space of physics, since the
scientist really makes use of causality to extend the intuitive spatio-temporal
order to regions not reached by perception. In this hierarchy the concept of
time comes neither before nor after that of space.

8. THIRD GROUP OF AXIOMS

III 1. Any effect of a given event which is not its limiting effect takes
place between two simultaneous limiting effects of this event.
III 2. Any event contemporaneous with event X has at least two limiting
effects which spatially coincide with X. Those events contem-
poraneous with X, which have one of these effects in common,
also have the others in common.
III 3. For event A, which is separated from event C by event B, to act
upon an event D, contemporaneous with C, it is necessary and
sufficient that it should act upon an event E which is contem-
poraneous with B and in turn acts upon D.
Axioms III 1-2 concern the properties of limiting effects. Translated into
optical language, the fIrst asserts the limiting character of the velocity of
light; the second amounts to saying that one can always send a light ray from
the place where X occurs which, reflected at the instant when X occurs, re-
turns to its point of origin. It goes without saying that this optical paraphrase
does not exhaust the meaning of these axioms. Axiom III 3 concerns the non-
existence of action at a distance in time. It could be roughly expressed by
saying that the past acts upon the future only through the intermediary of the
present.
THEOREM 6. The effects of an event A at instant t fills a sphere whose
PHYSICAL TIME 221

center coincides spatially with A. The spheres corresponding to two


simul-taneous events have equal radii.
Proof. If event A occurs at instant t, the sphere reduces to a point. There-
fore, according to Theorem 5, we will assume that A is not in t. According to
II 1, there is only one event, A', in t, which spatially coincides with A;
according to D 3 and III 1, there is in t at least one limiting effect, B, of event
A. According to D 6 and III 2, for event C to be located on the sphere with
radius A 'B and its center at A', it is necessary and sufficient that it should be
a limiting effect of A . An event located within the sphere always lies
between two events on the sphere, i.e., between two effects of A; according
to defmi-tion D 2, it also will be an effect of A. III 1 implies that there can be
no effects of A at instant t outside of the sphere. The second part of the
Theorem follows immediately from definition D 6.

THEOREM 7. An event which coincides spatially with event A is ar. effect


ofA. (Theorem 6).

THEOREM 8. The radius of the sphere of the effects of event A which are
contemporaneous with event B is equal to that of the sphere of the effects
of B contemporaneous with A.
Proof. Assume that A and B spatially coincide. Let C be a limiting effect
of A and contemporaneous with B, and let D be an event contemporaneous
with A and spatially coincident with C. The radius of the sphere of those
effects of A which are contemporaneous with B is equal to BC (Theorem 6),
that of the sphere of the effects of C which are contemporaneous with A is
equal to AD, and (according to the definition of congruence in enduring
space) we also have that AD = BC. The same will be true of the radius of the
sphere of the effects of B (or of any event simultaneous with C) since,
according to Theorem 6, this radius is the same for every event taking place
at the same instant. The hypothesis that B spatially coincides with A was
therefore not essential.

THEOREM 9. Two events which coincide with a third event in the space of a
given decomposition coincide with each other in the space.
Proof. Assume ftrst that event B, with which events A and C spatially
coincide, takes place in the interval of time separating them. Those effects of
A which are contemporaneous with B, fill a sphere around B in the instan-
taneous space containing B. The effects contemporaneous with C of events in
this sphere will fill a sphere which encloses all the spheres corresponding to
the events of the ftrst sphere in the instantaneous space containing C.
222 DURATION AND CAUSALITY

According to III 3, this enveloping sphere, having C as its center, is identical to


the sphere of those effects of A which are contemporaneous with C, whose center
spatially coincides with A, according to Theorem 6. Therefore, C occurs at the
same place as A. An analogous argument would apply to the case where B
occurred outside of events A and C in time.
A

If event B occurs in the time interval separating events A


and C, the radius of the sphere of the effects of A contemporaneous with
C is equal to the sum of the radius of the sphere of the effects ofA
contem-poraneous with B and the radius of the sphere of the effects of B
contem-poraneous with C.
Proof. Assume, in accordance with Theorem 9, that A, B, C, coincide in
space (this restriction is not essential, as in the proof of Theorem 8). The sphere
of the effects of A contemporaneous with C is the sphere enveloping the spheres
of effects corresponding to all the points of the sphere of the effects of A
contemporaneous with B. This enveloping sphere is obtained by tracing around
each point located within the sphere traced around C (with a radius equal to that
of the effects of A contemporaneous with B) a sphere whose radius is equal to
that of the sphere of the effects of B contemporane-ous with C. The radius of the
enveloping sphere is therefore equal to the sum of the two radii in question.

THEOREM 11. Let A and A', Band B', C and c', be six events which are
respectively simultaneous in the decomposition o. If B occurs between A
and C, B' does not occur outside ofA' and C'.
Proof. Designate by 'rCA, B)' the radius of the sphere of those effects of
A which are contemporaneous with B. Since B is between A and C in time, the
preceding theorem yields that rCA, C) > rCA, B) and rCA, C) > reB, C). If B'
occurred outside of A' and C', A' would be between B' and C', or C'
would be between A' and B'. In the first case it would mean that rCA, C) = r(A',
C') < reB', C') = reB, C), while in the second case, rCA, C) = r(A', C') < r(A',
B') = rCA, B), contrary to the hypothesis.

This theorem may be compared with Theorem 5, which guarantees that we


can order all the instants of a decomposition into a linear series, taking as our
point of departure a pair of events which are not causally related and which are
non-simultaneous in this decomposition; (Theorem 11) shows that the order thus
obtained is independent of the choice of this pair of events.
PHYSICAL TIME 223

9. THE TIME METRIC

Time considered as a magnitude rests essentially on the possibility of com-paring


time intervals, of deciding if the interval of time separating two instants of a
decomposition is smaller than, equal to, or greater than another interval, formed
by two instants of the same or of another decomposition. All of these questions
can be reduced to the simplest one, that concerning the equality of two time
intervals formed by instants belonging to the same decomposition. We accept the
following definition:
Let A, B, C, D, be four non-simultaneous events in a given decomposition.
We shall say that the time interval which separates A and B is equal to that
which separates C and D if the radius of the sphere of the events of event A
contemporaneous with event B is equal to the radius of the sphere of effects of
event C contemporaneous with event D [rCA, B) = r(C, D)]. (D 7)
Notice that the equality of time intervals has been defined relative to a given
decomposition. Two intervals which are equal in a certain decomposi-tion may
be unequal in another. Thus, the interval separating two events simultaneous in a
given decomposition will be the null interval in that de-composition, but will
have a fmite value in every decomposition in which these two events no longer
occur at the same instant. It is easily shown that the equality thus defmed satisfies
the following theorems:
THEOREM 12. The interval AB is equal to the interval AB, and also to the
interval BA (Theorem 8).
THEOREM 13. If interval AB is equal to interval CD, then CD is equal to AB.
(D?).
THEOREM 14. Two time intervals equal to a third are equal to each other
(Theorem 9).
THEOREM 15. If intervals AB and BC are equal to intervalsA' B' and B' C'
respectively, then AC is equal to A' C'. (Theorem 10).
THEOREM 16. If events A and B are simultaneous with event C and D
respectively, the interval AB is equal to the interval CD. (D 7).
This theorem brings out the fact that the equality of time intervals can be
extended to the instants of a given decomposition: the interval separating two
instants will be equal, by defmition, to the interval separating an arbitrary pair of
events which take place respectively at these instants. Theorems 12-15 apply also
to instants, which, moreover, satisfy the following propositions:
224 DURATION AND CAUSALITY

THEOREM 17. To each pair, t, t', of instants of a given decomposition, there


corresponds in this decomposition a single instant, til, such that tt" = t"t'.

Proof. Let A and A' be two events spatially coinciding in the decomposi-tion
{j,occurring at instants t and t' respectively of {j; let B be a limiting effect of A
contemporaneous with A'; fmally, let C be the center of the segment BA' (its
existence is guaranteed by II 4). According to III 2, event C has at least two
limiting effects which coincide spatially with A'. I say that one of these effects
must be between A and A'. We see first that there can not be more than one on
the same side of A, since this would contradict Theorem
10. It follows that C has only two limiting effects which coincide in space and
which are separated in time by A. Let us call D the one which is on the same side
of A' asA. We therefore have that A'D =A'C= CB, AA' =A'B, which by Theorem
10 entails that A'D = DA. Instant til is therefore the one which contains D. The
uniqueness of t" results from Theorem 10.
THEOREM 18. To each pair, tl and t2, of instant sofa given decomposition
there corresponds on each side of instant t3 of this decomposition a
single instant t4 such that tl t2 = t3t4 (II I, III 2).
These theorems show that time, when considered as a magnitude, has
properties analogous to those of a straight line. Shall we say that this analogy,
as well as defmition D 7, which derives temporal congruence from spatial
congruence, confirms Bergson's opinion that the measurability of time is only
indirect and rests essentially upon that of space? This appearance is only
illusory, since one could just as well define the measure of space in terms of
that of time, thereby demonstrating the quasi-temporal character of the
straight line. The defmition of the equality of two time intervals is a con-
vention whose connection with the measure of space is to a large degree
indeterminate.
This convention has often been discussed. 140 Poincare has maintained that, unlike
what occurs in the case of spatial equality, we have no intuition of temporal equality,
and that its definition is, therefore, completely conven-tional. On the other hand,
Enriques has insisted on the existence of rhythmic sensations, which do seem to
concern quantitative relations among intervals of time. Now, defmition (D 7) of
temporal equality is certainly conventional, as is that of spatial equality, but it is not
arbitrary. We require the spatio-temporal structure defined in terms of the causal
relation to be an indirect generalization of intuitive spatio-temporal data. The
generalized temporal scale must therefore contain the primitive scale as a special
case. This condition
PHYSICAL TIME 225

restricts, without altogether eliminating, the conventional character of the


defmition of the generalized scale, for the primitive scale can be generalized in
several ways - the choice of one of these ways is conventional. Thus, the
defmitions of temperature in terms of two different thermometric substances have
no priority with respect to each other, and the choice of one over the other can be
made only by convention. But both substances must meet specific conditions,
which result from primitive thermal data. A substance such as water, which does
not always expand with rising temperature, could not be chosen to define
temperature. The conventional element in definition D 7 is the completely
specific choice of the indirect generalization; what is no longer conventional is
the agreement between the temporal scale thus defined and the intuitive scale.
This defmition is theoretically expressible by means of the light ray clock. The
approximate agreement of this ideal clock with clocks in actual use is an
empirical fact resulting from the laws of physics; the isochronism of pendulum
oscillations, the deformations of springs, and the regularity of terrestrial rotations
are one consequence of well-established mechanical laws. The same is true of the
equality of the periods of a suppos-edly isolated cyclical mechanism. It is clear,
then, that by choosing another defmition of temporal congruence, by defming, for
example, the uniformity of time in terms of the rotating motion of the Earth, we
would shift the line separating convention and real knowledge in the physical
theory of time; the position of this line depends upon what axiomatization is
chosen. I believe, moreover, that what makes a scientist choose one defmition of
equality over another is not mere convenience, but rather a more rational
principle, namely that of determinism. If, instead of the variable t, which is
supposed to con-form with the principle of determinism, we were to introduce a
new variable t' = f(t) ((being a continuously increasing function which is
sufficiently close to unity so as not to run counter to intuitive data), the initial
conditions would no longer determine the evolution of an isolated system, since
these conditions also contain derivatives with respect to time, which would
change with the new defmition. Thus, among all the possible metrics which can
be attributed to time without changing the order of simultaneity and succession,
there is only one which satisfies the principle of determinism. The possibility of
choosing this metric is not conventional, but the choice of it, rather than of
another, is.

Notice also that in accepting the existence of the conventional factor in the
choice of a causal generalization of perceived time, we are by no means forced to
view the knowledge of this generalized order as in some sense more artificial or
more distant from reality than knowledge of the primitive order,
226 DURATION AND CAUSALITY

which is familiar to everyone and requires no new linguistic conventions


beyond those already existing in ordinary language. In fact, we have seen
that the knowledge of the local temporal order also rests on certain
conventions implicit in ordinary language. 141 We owe to Ajdukiewicz 142 a
general distinc-tion between the conventions added by science to ordinary
language (and recognized as such, especially since Poincare) and those
inherent in this language, the latter, epistemologically speaking, being no less
arbitrary than the former.

10. REFERENCE SYSTEMS, COORDINATES,


TRANSFORMATION FORMULAS
By a system of coordinates for a set we mean any one-one correspondence
between the elements of this set and the elements of a suitably delimited set
of numbers or groups of numbers. Every effectively definable system of
coordinates presupposes the use of a system of reference, i.e., a certain
number of individual elements serving as landmarks. Thus, when using or-
thogonal coordinates for Euclidean space, the reference system is usually
composed of three mutually perpendicular axes; in this case, the coordinates
are triplets of real numbers. A single reference system can give rise to several
coordinate systems, and vice versa, one system of coordinates may be at-
tached to several reference systems. These two concepts must therefore be
distinguished.
It is difficult to understand how a correspondence between real events and
numbers, which are extra-temporal and purely logical entities, can be established
effectively. We shall not discuss this problem here, but will limit ourselves to
examining how, starting from a group of events serving as a system of reference,
spatio-temporal coordinates are introduced into a causal decomposition. Let A
and B be two causally unrelated events, 5 a decomposi-tion in which A and B
occur at the same instant, and C and D two other events simultaneous with A and
B in 5, D not being located on the instantaneous plane formed by A, B, and C. We
have already said that such groups offour events uniquely determine a
decomposition: these groups will be called systems of reference. Notice that,
since the geometry is Euclidean at the instant at which event A occurs, to attach a
system of coordinates to this system of reference, we can construct a system of
orthogonal coordinates with A at its origin, the plane A, B, C, as the plane x, y, Z,
and an arbitrary segment (e.g. the segmentAB) as the unit oflength. All events
contemporane-ous with A will have the same temporal coor~inate, t = O. Event X,
which is
PHYSICAL TIME 227

not simultaneous with A, will have the same spatial coordinates, x, y, z, as event
X', which is contemporaneous with A and coincides in the space of 0 with X. The
temporal coordinate of X will have the same absolute value as the radius of the
sphere of those effects of X which are contemporaneous with A. To fIx the sign
of the temporal variable, we will adopt the convention of attributing to event A',
non-simultaneous with A, a fIxed sign. X will have the same sign as A' if it
occurs in the temporal interval separating A and A' or if event A' occurs in the
interval separating A and X. In every other case, Le., when A occurs between A '
and X, X will be opposite in sign fromA'.
It follows from the axioms and theorems cited above that this procedure
makes a unique group of real numbers, x, y, z, t, correspond to each event, and a
unique event correspond to each group of four numbers. If, instead of event A, we
chose another event simultaneous with it, or if we changed the spatial orientation
of the axes, the spatial coordinates would undergo an orthogonal transformation,
but the temporal coordinate x would remain unchanged. If we took as the origin
of the system an event which was non-simultaneous with A, for example, event
A', the t coordinate would be transformed according to the formula t = to + at',
to being the temporal coordinate of A' in the fIrst system and a being equal to
unity if the unit of length remains unchanged (Cf. Theorem 10). Finally, if the
two systems of coordinates were attached to two distinct decompositions,
their formula of transformation would be a Lorentz transformation.

To abbreviate the proof of this last statement we shall use the following
designations: a system of coordinates attached to a causal decomposition 0
will be called 'So' (this system is determined except for just two
substitutions: a linear substitution for the temporal variable, and an
orthogonal substitution for the spatial variables. We will say of such a system
that it is determined except for an auxiliary substitution). We will say,
furthermore, that a system of coordinates S (x, y, z, t,) determines a causal
decomposition, if the de-composition of all the events into classes, each of
which is defined by the condition that t = const., is a causal decomposition.
The assertion that the systems of coordinates attached to two such distinct
causal decompositions are inter-transformable by the Lorentz formulas is
therefore equivalent to the following two statements: (a) every system S'
obtained by a Lorentz trans-formation from a system S, attached to a causal
decomposition, determines a causal decomposition; (b) every system S
attached to a given causal decom-position 0 can be transformed by a Lorentz
transformation into a system S', attached to a given decomposition 0'. We
shall begin with the fIrst statement.
(a) Every Lorentz transformation applied to the system Sex, t, z, t)
228 DURATION AND CAUSALITY

attached to the causal decomposition 8 leads to a system S' (x', y', z', t'), attached
to a decomposition 8'. This means that: (1) two events having the same
coordinate t' are not causally related; (2) each event having the coor-dinate t'l has
at least one effect whose coordinate is t' 2 , provided that t'l = t' 2; (3) if events A
and C have the same coordinate t', every event B situated between them or
outside of them, in the sense of definition D 2, has the same coordinate, t'.

1-2. By using Theorems 6 and 10 and the definitions of the variables x, y, z,


t of the system S, it is easily shown that the necessary and sufficient condition
for the existence of a causal relation between supposedly distinct events EI(XI,
YI, zl> td and E2(X2, Y2, Z2, t2) is that their 'interval', determined by the
expression:
(Xl - X2)2 + (YI - Y2)2 + (ZI - Z2)2 - (ti - t2)2,
should not be positive. 143 But the Lorentz transformation:
, vt
=
X -
y' =y, Z' =Z, , t- vx Ivl< 1
X J"l-:::;2 , t = .,
J1=;2
leaves the interval invariant. The equality t'l = t' 2 therefore renders the interval
positive and entails the non-existence of the causal relation. It follows also that
every event E(X'2, y'2, Z'2, t'2) having the coordinate t'2 =1= t'l is causally
related to the event E(x' 1, y' I, Z' 1, t'l), provided that its coordinates satisfy the
following inequality:
(X'I -X'2)2 +(y'l -y'2)2 + (Z'l -Z'2)2 -(t'l -t'2)2 ~O

3. Thanks to an appropriate auxiliary transformation, we shall be able to


assume that the events A', B', and C', simultaneous in 8 and coinciding
respectively in the space of 8 with the events A, B, and C are located in the plane
xy of S6, that the x axis passes through A'C' (in that order), and that the origin of
S is at A. It follows that the effects of A, B, and C having the same coordinate, t',
fill the spheres defined respectively by the following inequalities:

(A) X'2 + y'2 + Z'2 ~ t,2


(B) (x'-X'B)2 + (y'-y'B)2 +Z'2 ~(t'_t'B)2.
(C) (x' -x'd2 + y'2 + Z'2 ~ t'2
If B occurs between A and C, i.e., if each effect common to A and C is also an
effect of B, then each group (x', y', z', t') satisfying the inequalities (A) and (C)
must also satisfy inequality (B). But this is possible only on the
PHYSICAL TIME 229

condition that t'B = O. More precisely, the necessary and sufficient condition for
each group (x', y', Z', t ') satisfying inequalities (A) and (C) to satisfy inequality
(B) is that:
(D) O<X'B <X'c, Y'B = t'B + 0.144
The condition is sufficient. For x' ~ X'B we have that (x' - X'B)2 < (x' - X'c)2,
therefore, (C) implies (B). For x' > X'B we have that (x' - X'B)2
< X'2, therefore, (A) implies (B). Consequently, the simultaneous validity of
(A), (C), and (D) always entails that of (B).
The condition is necessary, for, if we assume at least one of the relations
(D) to be unsatisfied, we can always find a point which satisfies (A) and (C)
but not (B): the point (iX' c J ± J t'2 - tx'2 C , 0, t') satisfies (A) and (C) but does
not satisfy (B) for a suitable value of t ' and a suitable sign for y', unless all the
relations (D) are simultaneously satisfied.
We conclude from this that if event B is situated between two events A and C
which have coordinate t, B will also have that coordinate. We shall now consider
the case where B occurs outside of A and C, assuming, to fix our ideas, that A
occurs between Band C. This case can be reduced to the pre-ceding one with the
help of the follOWing two easily proved Lemmas : (a) the existence of event Z
between events X and Y entails the non-existence of a causal relation between X
and y;14s ({3) if there is no causal relation between events U and T, we can
always transform So into S =S{, so as to obtain tu =
Since A is between Band C, we conclude from (a) that there is no
causal relation between Band C, and from ((3) that we can, by a Lorentz
transformation, pass from S to a system Sex, z, ji, 1) in which lB = lC' From
this and the preceding case it follows that lA = lB = tc, which implies that tA -
VXA = tB - VXB = tc - vxc. The relation among the numbers tA, tB, tc, XA,
XB, xc, drawn from these equations by eliminating ii proves that the events A,
B, and C have the same temporal coordinate in every system obtained from the
system S by a Lorentz transformation, which renders eventsA and C
simultaneous, hence, in particular, in S'.
We have just shown that by affecting a Lorentz transformation on a system of
coordinates attached to a given causal decomposition, we always obtain a system
of coordinates attached to another causal decomposition. It only remains to be
shown that we can thus obtain every system of coordinates attached to causal
decompositions, i.e., that we can always pass by a Lorentz transformation from a
given system So to a given system So '. Let us consider two events, A and B,
occurring at instant t of O. If they are not simultaneous in S' = So', Lemma ({3)
tells us that we can fmd a Lorentz transformation
230 DURATION AND CAUSALITY

, leading from the system S' to a system S" in which A and B are simultaneous.
All events located on the straight line determined by A and B are also simul-
taneous in Sand S". Consider an event C occurring at instant t but not this
line. If this event is not simultaneous with A and B in S", there will be a
Lorentz transformation leading from S" to a system S'" in which A, B, and C
are all simultaneous. (By taking the straight line through A and B as the z"-
axis, and by taking into account the non-existence of a causal relation
<
XC"2bet
weenCandanarbitraryeventonthisline,weobtainthattc"2
+ YC,2, i.e., that event C occurs outside the cylinder of radius tc" drawn
about the axis z". If we eliminate YC" by an auxiliary transformation, we
obtain Itc" 1< Ixc"l, and if we define S'" by the parameter v = ~~':, , we
will obtain tA '" = tB'" = tc'''). Finally, let us choose event D which occurs at
instant t but not on the plane determined by A, B, and C. If D is not
simultaneous with A, B, and C in S'" we may still, reasoning in the same
manner, find a system S"" in which the four events, A, B, C, and D, are all
simultaneous. We conclude from this that the decomposition 0 and 0"" (the
latter dermed by the equation S"" = Sfj"") have an instant in common and are
consequently identical, by Theorem 20. 147 The resulting transforma-tion
which leads directly from S' to S"" therefore satisfies the condition in
question: it leads from the given system Sfj' to the system Sfj.

11. INERTIAL SYSTEMS

Up until now we have considered the spatio-temporal coincidence of two


events, defined by the community of their effects, as synonymous with
numerical identity. Although this view simplifies language by permitting the
expression 'set of events coinciding with a given event' to be replaced every-
where with the term 'event', it is by no means necessary for the coherence of
the axioms. We shall now abandon this simplifactory convention, since the
plurality of events located at the same world point seems to be the most
fundamental characteristic of physical reality, insofar as the latter transcends
any purely geometrical structure. Space-time viewed as a set of world points
is homogeneous: every geometrical property belonging to a world point
belongs to every other as well; it is impossible to define in geometrical terms
an absolute length which would have characteristic properties. However, all
this is completely changed when we distinguish the event from the world
point. The fortuitous, unpredictable character of the physical universe then
makes its appearance.
Physics distinguishes two sorts of fundamental entities, matter and field.
PHYSICAL TIME 231

The most elementary form of matter is the elementary particle. In discussing


the concept of duration 148 we have tried to show that this identity through
time cannot be interpreted as numerical identity, especially if we take the
relativity of time into account. Each enduring particle must be conceived as a
set of instantaneous particles. The converse of this proposition is not true,
however. We have described the additional condition which must be met by a
set of instantaneous particles for it to be considered as an enduring particle
- i.e. the double postulate of continuity in time and discontinuity in space.
Thus, the identity of an electron through an interval of time consists in the
fact that certain physical properties of space-time, like the distribution of
density, change in a continuous fashion along the world line of the electron
while changing discontinuously in its spatial vicinity. By taking account of
the fact that the presence of a particle at a world point means the presence of
one more event in the set of events constituting this point, we can now
characterize the double postulate of temporal continuity and spatial dis-
continuity much more simply in terms of our axiomatic system. Under the
simplest hypothesis - the only one we shall discuss here - in which the only
events under consideration are the instantaneous particle itself and the
instantaneous particle of the magnetic field, we can say quite simply that a
world point containing only one event is, by definition, devoid of matter,
while a world point containing two events is supposed to contain an instan-
taneous particle. An enduring material particle, by definition, will be a
continuous world line which is composed of world points containing two
events and which traverses a region each of whose world points contains only
one event. We will say that a given material particle is at rest in a given
decomposition if all the instantaneous material particles which it contains
coincide in the space of this decomposition. A material particle at rest in one
decomposition is in uniform rectilinear motion in every other decomposition.
Accelerated points are not at rest in any decomposition. A set of material
particles, all of which are at rest in the same decomposition and whose instan-
taneous points occupy a linear segment at each instant of this decomposition,
is a rigid rod at rest in this decomposition. A set of three mutually perpen-
dicular rigid rods at rest in a given decomposition is an inertial system. The
inertial system is therefore a superabundant frame of reference: in place of
four events, it contains a two-fold infinity.
It is, however, these infinitely superabundant frames of reference which
predominate in practice. By a frame of reference, the physicist understands a set
of three rigid rods, and not a group of four suitably delimited elemen-tary events.
When we say, for example, that the trihedron formed by three
232 DURATION AND CAUSALITY

contiguous walls of a room constitutes a system of reference for an observer


at rest with respect to it,149 we seem to attribute to this observer the ideal
faculty of being capable of assigning to each event (no matter how far in the
future or the past) the four coordinates describing its spatio-temporal orienta-
tion with respect to the system of reference. But, if the observer had this ideal
faculty, it would be necessary for him to take as his point of departure not the
knowledge of the whole trihedron through time, but rather that of four
simultaneous events at four points of the trihedron, for example, that of four
tiny instantaneous spots which the light of the sun had simultaneously
projected onto the edges of the trihedron. In reality, this ideal faculty is very
limited. The memory of the observer is too weak to retain for any length of
time the image of the four instantaneous spots and the places where they
appeared. At the least, he would need four spots which endured and were
visibly at rest with respect to one another. But even this would be too in-
convenient: even supposing that we had succeeded in obtaining four colored
spots which agreed not to move, it would still be very difficult to mark the
position of an event, however imprecisely, with respect to these spots. We
would therefore draw out the spots into sensibly stationary lines, which it
would be to our advantage to make mutually perpendicular. For example,
they might be three fme jets of some colored liquid, emerging at right angles
from a suitably constructed fountain. But the upkeep of such a reference
system would be very inconvenient. We will therefore replace the liquid with
a solid, to form three easily preserved bars, and then risk the hypothesis that
these three solid rods constitute a system at rest in a causal decomposition, an
inertial system.
This hypothesis concerns the relation of the three rods to the set of all
events, but it entails consequences verifiable by local observations, mechani-
calor optical, such as the test of the three stones proposed by Lange, ISO or the
optical experiments described by Reichenbach. 1s1 In these observations the
spatio-temporal order is directly given; all the relations of simultaneity,
succession, coincidence, and spatial congruence, when linked to the decom-
position in which the observer is at rest, are directly (although only approxi-
mately) accessible to perception. The addition of events not accessible to
perception, the extension of the system to distant regions of the universe,
will then be done by causal methods. We see, then, that the use of rigid
inertial systems is not inherent to the spatio-temporal structure of becoming,
but rather that it is a practically unavoidable means of linking the perceived
spatio-temporal order to the physical representation of the universe.
Furthermore, Axioms I-III, which deal with the geometrical properties of
PHYSICAL TIME 233

space-time, are hypotheses, just as much as the propositions asserting the


existence of particles and inertial systems. If these geometrical properties of
space-time seem to us to be less contingent and more rational than the
physical properties attached to the existence of matter, it is because they have
the singular merit of occurring in a simple system of axioms, while the
distribution of matter at a particular instant seems to be a basic fact, which
will never be known even imprecisely, and which will never be given by any
formula. It is clear, moreover, that Axioms I-III do not imply even the
existence of a single elementary particle, since these axioms would also be
satisfied if each world point were composed of only one event. If follows a
fortiori that we could not deduce the particular distribution of world lines, which
corresponds to the atomic structure of matter, from these axioms. The existence
of an extended particle, such as an electron or a proton, corresponds in space-
time to a four-dimensional tube with temporal extension. The fact that all
particles of the same kind (all electrons, for example) have the same properties,
is expressed by the identity of the diameters of instantaneous cross-sections of
these tubes, with respect to decompositions in which the particles are at rest. We
can see, moreover, that our criterion of identification does not permit us to follow
the history of a particular point of the electron in time since the postulate of
spatial discontinuity is not satisfied in this case; according to this criterion, only
the electron conceived as a whole endures, while the points of which it is
composed have no individuality of their own.
The fact that the existence of discontinuous matter entails that of privileged
lengths, definable in terms of the causal relation (for example, the maximum of
minimum length of the diameter of a four-dimensional tube of discontinu-ity),
seems important from the epistemological point of view. In physics there are no
rigorously verifiable laws or magnitudes measurable with absolute precision.
Each physical computation involves some 'negligible' quantities. All the
fundamental physical concepts imply the possibility of 'neglecting' something: in
the elementary particle its dimensions are neglected, in inertial motion,
acceleration, in isolated systems, the action of external forces, in electricity, the
action of small charges said to explore a field, in thermo-dynamics, the duration
of the process. Physicists all agree that neither ele-mentary particles nor inertial
motions, nor any other simplified entities exist in reality. The possibility of
basing a science of the real on these fictions rests on the fact that all the
complications, which are assumed in the theory to be non-existent, are in reality
only negligible. But negligible with respect to what? In homogeneous space-time,
nothing is negligible. Each segment, however small, is very large with respect to
a smaller segment. The existence
234 DURATION AND CAUSALITY

of discontinuous matter implies the existence of absolute units of length.


'Negligible' in a given problem means 'of the order of one of these absolute
magnitudes'. The possibility of making these approximations, which are
characteristic of the physical knowledge of the real, seems to be linked to the
existence of discontinuous matter.

12. OPTICAL CRITERIA

Since it transmits a local disturbance enclosed in a surface of discontinuity, a


light ray can also be characterized 152 by the double postulate of continuity in
time and discontinuity in space. The essential difference between a material
signal and a light signal is that only the latter travels with the maximal velocity,
(or in the terms of our axiomatic system, each ofthe two events linked by a light
signal is a limiting effect of the other). From this statement we derive the optical
criterion of simultaneity utilized by Einstein.

If a light ray joining events A and C coincides in Cwith a ray


joining Band C, and if events A and B coincide in the space of a
decomposi-tion 8, in this decomposition C will be simultaneous with
event D, which marks the center of the temporal segment AB.
Proof. Event C is a limiting effect of both A and B. Let D be the event
which is simultaneous with C and spatially coincides with A. According to
Theorem 6, segment CD will be equal to the radius of the sphere of those effects
of B contemporaneous with C (D, which spatially coincides with A, also
coincides with B, by Theorem 9). By defmition D 7, the temporal distances DA
and DB are equal. Therefore, the center of the segment AB is simultaneous with
event C.
Notice that spatial coincidence, assumed in this criterion to be known,
concerns the events A and B, which can be close together in time, while, thanks
to the enormous speed of, light, C can be very distant in space. For example, if
we assume that A and B are one second apart (thus rendering their spatial
coincidence easily perceivable) event C can be as far as 300,000 kilo-meters
away from them. However, what makes the criterion theoretically important is
the fact that the decomposition 8 is completely determined, as the next theorem
will show, by the two events,A, B, which spatially coincide in it; the entire
spatio-temporalstructure seems, therefore, to be defmable in terms of a single
pair of perceivable events.
THEOREM 20. Two decompositions are identical if there exist two events
which spatially coincide in both.
PHYSICAL TIME 235

Proof. Assume that events A and B coincide spatially in decompositions {j


and l)'. By D 5, there is in [) an instant t in which the spheres of the effects of
A and of B coincide, forming only one. Consider two events, C and D, located at
the surface of this sphere. From Theorem 6 we get that each is a limiting
effect of both A and B, which, according to Theorem 19, entails that C and
D are also simultaneous in [)'. All the events located in this spherical surface
are therefore simultaneous in both decompositions. The same is true of all
other events occurring at instant t, since each of them can be reached by a
straight line having two points in common with this spherical surface (it fol-
lows from defmition D 3 that any event located on the straight line
determined by events X and Y, simultaneous in a given decomposition, is
simultaneous with X and Y in every decomposition in which X and Yare
simultaneous with each other). {j and l)' will therefore have instant t in
common. 15 3 But two decompositions having an instant in common are
identical. Let us assume that at instant tl of {j there are in fact two events, X
and Y, which are not simultaneous in {j'. According to I 3, the straight line
passing through X and Y would have a point in common with each instant of [)',
and therefore in particular with t, which is impossible, since, being located in
instant t 1 , it can have no event in common with instant t, two instants of the
same de-composition never having an element in common.
THEOREM 21. Two decompositions are identical if there are four events
simultaneous in both, but which are not all located in the same plane.
These decompositions will then have an instant in common. The fact that
four events coinciding in time are necessary to determine a decomposition,
while only two events coinciding in space are needed, shows that spatial
coincidence is a closer connection than simultaneity. The latter presupposes
the equality of three coordinates, the former concerns only one. From the
point of view of local order, spatial coincidence seems just as fundamental as
temporal coincidence; either can serve as the point of departure for the
causal construction of space-time.
Finally it is important to notice that in the formulation of the optical criterion
of simultaneity, no use is made of the temporal orientation of the light signal:
events A and C are said to be linked by a light signal regardless of which of them
coincides with the departure or arrival of the ray. This is because simultaneity is a
symmetrical relation whose determination does not depend upon the choice of
temporal orientation. But would it not be possible to find a criterion for
asymmetrical succession among the properties of light signals? This question can
be understood in two ways. On the one hand, using
236 DURATION AND CAUSALITY

local observations, we could try to determine if a particular passage of a light ray


is a 'departure' or an 'arrival'; thus, a ray produced at a point by the striking of a
match would be taken as departing from that point. We have already discussed
this criterion, which derives the local anisotropy of physical time from a
fragment of psychological time but does not imply the universal anisotropy of the
former. On the other hand, however, we might try to define an intrinsic
anisotropy of physical time by using optical signals. This seems to be the opinion
of Weyl154 who sees the criterion of an absolute difference between the physicai
past and the physical future in the mode of propagation of light waves, since
these waves, although theoretically divergent or con-vergent, in fact always seem
to diverge. This hypothesis may be expressed geometrically by saying that light
rays, which have a beginning and no end, form half-lines all oriented in the same
temporal direction, which, by de-finition, will be the direction of the future. It is
clear that if we add this supplementary hypothesis to axioms I-III we will be able
to define an asymmetrical succession which would be linked to the isotropic
temporal order by the following proposition: 'If event A occurs before event B,
which in turn occurs before event C, event B occurs in the interval of time
separating events A and C'. I believe, however, that this new axiom, which would
clearly be independent of the axioms I-III, would add nothing essential to the
spatio-temporal structure defmed by them. [In fact, the axiom was refuted by a
remark Einstein made in 1910. Cf. Part III, H. M.J
x

NON-PHYSICAL TIME

1. PSYCHOPHYSICAL TIME: THE PERCEPTUAL METHOD

In practical life and in science, we employ several methods of


psychophysical synchronization. We shall describe these methods, which are
unconsciously applied in the most banal situation, because, if we wish to
understand the psychophysical temporal order, we must first establish how
this order is effectively recognized.
If upon feeling a toothache and noticing at the same time that the two hands
of the clock coincide with the numeral twelve, I claim to have been in pain at
noon: I have established a simultaneity between the psychological state of the
pain and the physical event of the coincidence of the clock's hands, i.e., a
psychophysical simultaneity. How can we distinguish the im-mediate datum from
the hypothetical factor in such a judgment? What is immediately given is the
intuitive simultaneity between the sensation of pain and the visual image of the
coincidence of the hands, while the fact that this visual image of a coincidence
corresponds to a real coincidence of the hands is only hypothetical.

This distinction is independent of the choice between the idealist or the


realist point of view. Under a realist interpretation, the real coincidence of the
hands is a fact distinct from the image of the coincidence and does not
necessarily accompany the latter. On the idealist interpretation, the
coincidence is not a reality independent of any perceptual image, since it
resides in a special interconnection of perceptual images ~ that of the given
image's belonging to a determinate system of images. But this system of
images is still distinct from the given image, the perception of which does
not guarantee that it will be part of the system. The fact that the image of the
coincidence of the hands is accompanied by their real coincidence is
therefore equally as hypothetical from the idealist point of view. Thus, the
psychophysical simultaneity in the example cited consists in the following
combination of the perceived and hypothetical factors: a psychological state
is psychophysically simultaneous with a physical event if it is intuitively
simultaneous with a perception of this event.
This is the fundamental rule which we use to synchronize our own psycho-
237
238 DURATION AND CAUSALITY

logical states with the events of the material world. By associating the events of
our inner life with objective dates, determined by the clock and the calen-dar, we
use the relation between the perception and the perceived event as the uniting
factor between psychological time and physical time. But the same method may
also be used to synchronize the psychic life of others with the material reality.
The historian often has recourse to it in establishing the dates of some historical
personage by means of a physical event which this personage is supposed to have
witnessed. It is in this way that a solar eclipse mentioned by a Greek philosopher
can be used as a reference point for his biography.

The first method of synchronization, which we shall call perceptual, is,


moreover, only approximate. In fact, in the example cited above, the coin-cidence
of the hands is not rigorously simultaneous with its perception, but imperceptibly
prior to it. If, instead oflooking at a nearby clock I listened to the striking of a
distant one, I would no longer be able to apply this method without arriving at
unacceptable consequences (it could happen, for example, that two physical
events synchronized by this method to the same psycho-logical state are not
physically simultaneous with each other). This is because this method takes into
account neither the time of propagation of the physical stimulus nor the time
elapsed between the stimulation of the sensory organ and the corresponding
central disturbance: the perception of the coincidence of the hands is only
indirectly linked to this coincidence, through a temporally extended process
determined by this coincidence whose principal phases are the incidence of a
light rayon the hands at the moment of their coincidence, its reflection, the
traveling of the wave or the photon toward the eye, the stimulation of the optic
nerve, and finally, the disturbance of the optic center. In this complex process,
which is inserted into physical becoming without disturbing its continuity (since
the coincidence of the hands with which it begins is no more than the propagation
of their prior movement, and since the disturbance of the optic center, where it
terminates, is subjected in its turn to further transformations), perception
delineates two quasi-instantane-ous phases, its physical object and its
physiological substratum, in other words, the coincidence of the hands and the
fmal disturbance of the optic center. We will not analyze here the reasons for
which we select, from the infinite set of phases which constitute the physical
process corresponding to a given perception, a completely determinate pair of
these phases as being the physical object and the substratum of the psychological
state, nor, in particular, whether the mysterious correspondence, which links the
visual sensation of the coincidence with the final perturbation of the optic center,
NON-PHYSICAL TIME 239

signifies identity, parallelism, interaction, or simple hitching, as Bergson


thinks. The approximate value of the perceptual method of psychophysical
synchronization, and the temporal difference which exists between the
departure of the physical stimulus and the central perturbation, show that it is
this mysterious correspondence between the psychological state and the
cerebral state which creates a time common to the psychological and
physical orders: if the coincidence is anterior (in the psychophysical sense) to
its perception, this is because it is physically anterior to the physiological
sub-stratum of this perception, i.e. to the perturbation of the optic center. We
propose to consider this statement as a definition of psychophysical succes-
sion, and to assume, consequently, that in order for a physical event X to be
anterior to a psychological state Y, it is necessary and sufficient for event X to
be physically anterior to the physiological substratum of state Y. It will
doubtless be objected that it is often possible to determine the succession of a
physical event and a psychological state even though the physiological
substratum of the psychological state is unknown. However, a closer
examina-tion shows that the psychophysiological relation is always implied
in the determination of a psychophysical succession, at least when the latter
is effected by means of the perceptual method.
If I believe the coincidence of the hands is earlier than its perception, even
though I am incapable of indicating the physiological substratum of the latter,
it is because this coincidence is also earlier than the stimulation of the retina,
which in turn precedes the disturbance in the optic center. Thus, thanks to the
transitivity of physical succession, I can consider the stimulation of the
retina, or any other easily observable phase occurring between the
coincidence and the fmal disturbance instead of the disturbance in the optic
center; but the very choice of an intermediate phase between the coincidence
and the physiological substratum of the perception shows that the latter is
implied.
Let us point out the causal character of this definition of psychophysical
succession. In saying that it is the final disturbance of the optic center which
is the substratum of the visual perception, we do not wish to assert that this
disturbance is not followed by other disturbances of the same center - this
would obviously be false - but only that it is the last disturbance of this
center on which the perception depends, later disturbances having no
influence on this perception. This amounts to saying that since the sensory
process, beginning with the coincidence of the hands and leading to the
disturbance of the optic center, is not isolated from what takes place in the
organism and its immediately vicinity, we can, by changing these 'boundary
conditions',
240 DURATION AND CAUSALITY

change the process itself, and we will notice that such a modification of the
sensory process is followed by a change in the perception, provided that the
change in the boundary conditions takes place before the physiological
substratum of the perception.
The isotropic character of the temporal order defined by the psycho-
physical relation results from the fact that this relation essentially plays the
role of simultaneity in the definition and therefore cannot introduce an
anisotropic factor into the temporal order, since the concept of simultaneity
does not depend on which of the two opposite temporal directions we might
choose. This is also illustrated by the following remark: the statement that
the sensory process, which begins with the coincidence of the hands, leads to
the disturbance of the optic center and is always accompanied by the per-
ception of this coincidence, introduces no difference between the beginning
and the end of the process, since it is to the process as a whole that the
perception corresponds.
The question then arises of how a temporal location is to be assigned to the
perception with respect to the associated process. The definition of
psychophysical simultaneity in terms of the psychophysiological relation
resolves this question in the following manner: in the process beginning with the
coincidence of the hands, to which the perception of this coincidence is
associated, there is a unique phase rp such that the change of any phase
contained in the interval between rp and the coincidence of the hands involves a
change in the associated perception, while changing a phase of the process
outside this interval brings no change in the perception. This unique phase which
separates the two kinds of phases of the sensory process, and is called the
physiological substratum of the perception is, by definition, simultaneous with
the perception. We see that in this explanation no use is made of before and after;
all the relations utilized assume only an isotropic time.

2. PSYCHOPHYSICAL TIME. THE EXPRESSIVE METHOD

The perceptual method, however, leads itself directly only to synchronization of


physical events with psychological states which are perceived simultaneous-ly
with these events; it becomes inapplicable when it comes to our past states and
especially to the psychological states of others, which, in general, are accessible
to us only through the bodily processes which accompany them. It is through
these accompanying processes that the psychophysical synchroniza-tion, which
goes beyond the present perceived by the observer, is generally effected. If, while
noticing that both hands of a clock cover the numeral
NON-PHYSICAL TIME 241

twelve, I perceived someone's face contorted with pain, I would be


practically certain that this person suffered at noon. Which are the intuitive
and inferred factors in this psychophysical simultaneity? The psychological
simultaneity of the perceptions of the coincidence of the hands and the facial
expression is immediately given. What is hypothetical is whether (1) the
perception of the coincidence of the hands and the facial expression
corresponds to a coincidence and an expression which really exist; and
whether (2) this facial expression in fact corresponds to a state of suffering in
this person. The fIrst of these questions also appears in the perceptual
method. It is the second which is characteristic of the type of synchronization
which we call 'expres-sive', and which can be summarized in the following
statement: 'A physical event is simultaneous (in the psychological sense) with
a psychological state if it is simultaneous (in the physical sense) with another
physical event which expresses this psychological state'.
It is obviously this second rule which governs our knowledge of the temporal
order of the psychological states of others. If we say of someone that he is now
hungry or ashamed, gay or sad, it is because we now see him eat or blush, laugh
or cry. The presence of these bodily motions leads us to believe in the presence
of the psychological states which they express.
The expressive and perceptual methods of psychophysical
synchronization both seem to be practically indispensable. A correct
definition of psycho-physical simultaneity must therefore account for them
both. I believe that this is true of the defmition put forth on pages 238-239,
since the expressive method is based upon the same psychophysiological
principle and has the same causal character as the perceptual method. In
order to show this, we will examine more carefully the particular
psychophysical relation on which the expressive method is based.
This relation, which consists in the correlation of a particular physical
phenomenon, e.g. the act of blushing, with some psychological phenomenon,
e.g. the feeling of shame, which it is taken to 'express', has been the object of
much study by psychologists, particularly since Darwin's well-known
work.Iss It has become possible to characterize several psychological states
precisely in terms of the bodily phenomena which accompany them. More-
over, here as elsewhere, scientific research has only extended the primitive
sciences of common sense and art. We all know in practice how to interpret
psychologically the expressions and gestures of others, and the descriptions
of psychological states attempted in the plastic arts and in literature are also
necessarily limited to these bodily phenomena. What interests us here is to
determine precisely what it means to say that certain psychological states
242 DURATION AND CAUSALITY

have a tendency to find an expression in a bodily phenomenon. This problem is


analogous to that of the causal relation. The sciences establish a multitude of
relations of causality among particular phenomena; the analysis of the causal
relation in abstracto is a philosophical problem. This analogy between the causal
relation and the expressive relation is not merely formal: we can define both, to a
fIrst approximation, in terms of regular succession. If physical event A is
regularly followed by physical event B, B is considered as an effect of A; if a
psychological state X is regularly followed by the bodily state Y, Y is an
expression of X. Paleness is an expression of fear, because it regularly
accompanies it. 15 6
The following argument will perhaps be urged against this definition of the
expressive relation: if every physical event B, which regularly follows the
psychological state A, is an expression of A, and, if every physical event C
regularly succeeding B is an effect of B, then every effect of a physical event
which ,expresses a given psychological state is also an expression of this state.
The expressive function of a bodily phenomenon would therefore be trans-mitted
by the relation of cause and effect. But this conclusion does not seem to be
always verifIed, Undoubtedly, the words recorded on a record may be said to
express the thoughts of the speaker, since they are linked by a causal relation to
the words which he pronounced and which directly expressed his thought.
However, the words of the speaker are also causally related to other physical
phenomena, for example, to the imperceptible vibrations of the walls of the room
where they were pronounced, without these vibrations, produced by sound waves,
being considered as expressing his thought.
This difficulty will be removed if we consider another analogy between the
expressive relation and the causal relation. We distinguish, especially after
Schopenhauer, the total cause from partial causes. If the complex event A is
always followed by event B, we call A the total cause of B, while any event
forming a part of A is a partial cause of B. Thus, when the fall of a spark on a
powder keg sets off an explosion, the total cause of the explosion includes not
only the fall of the spark (which, in itself, would not suffice to cause the
explosion), but also the instantaneous state of the powder keg and of its
immediate environmnet at the instant in question. The fall of the spark is only
one of the partial causes, although practically the most important since, having
come last, it set the process in motion. But this fact is of no theoretical
importance, for the other determining conditions of the explosion might also
have been produced at the last moment. Similarly, we will distinguish the total
effect of a given cause from its partial effects, the total effect being defmed as the
set of partial effects realized at a given time (and not as the set
NON-PHYSICAL TIME 243

of all the partial effects, just as the total cause of an event is not the set of all
its partial causes, but only the set of partial causes realized at a given time). 15
7 The causal relation linking an arbitrary partial cause to some partial effect is
obviously not a one-one relation, since each cause has many effects and each
effect many causes. However, this relation becomes one-one if it applies to a
total cause and a total effect, while taking their locations in time into
account: every event has only one total cause at a given time in its past, and
only one total effect at a given time in its future.
It seems to us that the same distinction must be made between the total
expression and partial expressions of a given psychological state. When we say,
for example, that the scream expresses a sudden fright, this expression can be
only partial. The actor who screams on the stage often experiences no fright; the
scream, which generally expresses fright, no longer does so with him. In what
does the difference consist? Just as we cannot conclude with certainty the
existence of the effect from the realization of a partial cause, similarly the partial
expression of fright, which is the scream, does not permit us to conclude with
certainty the realization of the correspondhg psychologi-cal state. 'In the actor the
scream does not express fright' means therefore that in the organism of the actor,
phenomena do not exist which are simultaneous with this scream, and which, as a
set, completed with the phenomenon of the scream, are regularly preceded by a
fright (although, in general, the phenomenon of the scream is accompanied in the
organism of the crier by a group of phenomena with which it constitutes a
complete expression of fright). These additional phenomena undoubtedly include
specific vasomotor and respiratory changes.

It is this difference between the total and partial expressions which also
explains why the effect of a bodily gesture does not always seem to us to express
the same psychological state as the primitive gesture. I do not believe we would
hesitate to consider the total effect of a gesture as expressing the same
psychological state as the gesture itself, since, given the one-to-one relation of
the total effect to the total cause, the knowledge of the total effect allows us to
reconstruct (at least in theory) the primitive expression as precisely as we wish
and thereby to go back to the same psychological state. For example, if a
psychological state is expressed by an impulse given to a thereafter isolated
material system, all the subsequent states of this system must be considered as
expressions of this psychological state, just as much as its initial state, since,
from the knowledge of one of them, we can always ideally reconstruct the initial
state, and therefore the psychological state as well. Thus, an isolated painting
would not cease to express ideally the painter's
244 DURATION AND CAUSALITY

VISIon even if it kept 'deteriorating' as time went by, since, assuming the
painting to be isolated from external influences and subject only to internal
forces, we can, from a knowledge of its fInal state, reconstruct its initial
state, and through it, attain the artist's vision. This will no longer be true if we
are dealing with a partial effect, where the painting was not isolated. In this
case, its initial state is no longer explainable from its fInal state alone, which
will have to be complemented by external influences; a knowledge of it will
not suffIce to lead us to the psychological state of the artist. It is therefore
only in the case of a transition from one cause to its total effect that this
effect continues to completely express the psychological state expressed by
the cause. The partial effect of a physical process expressing a given
psychological state will be only a partial expression of this state.
Is the expressive function also preserved in the inverse direction, from
effect to cause? This appears to follow from the fact that every effect is
completely determined by its total cause; therefore, if the effect is suffIcient
to determine the psychological state, its cause will also be suffIcient. This
conclusion is verifIed when we pass from a physical phenomenon external to
the organism of the person, whose psychological state this phenomenon
expresses, to a phenomenon taking place at the surface of this organism.
However, one would hesitate to consider conditions prior to a phenomenon
expressing this state as expressing the same psychological state if these con·
ditions are found inside the organism: we might consider a change of facial
color as expressing certain affective states, but would we say the same of the
accelerated heart·beat which conditions this change, (even though from the
introspective point of view, the two bodily phenomena seem to be equally
essential)? I think that this hesitation would disappear if we could observe
someone's heart·beat for a suffIciently long time (e.g., by means of a cardio·
graph) and if we thereby learned to recognize the expressive function of that
person's cardiac movements. Moreover, these cardiac movements are
conditioned in tum by certain changes in the central nervous system. Some·
one aware of the causal dependence between the accelerated heart·beat and
the corresponding state of the central nervous system could surmise the
psychological state of the person under observation starting from the state of
the nervous system, just as well as from the cardiac phenomenon, and if he
considered the state of the heart to express the psychological state, he would
have to do likewise with the nervous system. Therefore, the property of
expressing a given psychological state, which was attributed to a momentary
facial change, is preserved in passing from this superfIcial change to the
cardiac phenomenon which conditions it, and, from this to the cerebral
NON-PHYSICAL TIME 245

phenomenon conditioning it in turn. But it is clear that this descent into the past
cannot be continued indefinitely. In passing from a nervous state to an earlier
state always expressing the same psychological phenomenon, we will finally
reach a nervous state simultaneous with this phenomenon, i.e., its physiological
substratum. The nervous states prior to this substratum (and, consequently, to the
psychological phenomenon) could not be said to express the phenomenon, since
a psychological phenomenon cannot be expressed before it itself takes place.
Therefore, a bodily phenomenon expresses a psychological state if it is
an effect of the physiological substratum of this state.

This statement establishes an interconnection among three relations: causal,


psychophysiological, and expressive. If we take two of these relations as given,
we can interpret this statement as defining the third in terms of the other two,
provided that the interconnection can be resolved with respect to
this third relation. Thus, the psychophysiological relation can be conceived as
linking the psych~logical state to its initial expression - in other words,
among all the bodily phenomena expressing a given psychological state, that one
which precedes all the others in time is taken to be the physiological substratum
of this psychological state. This interpretation would allow us to define the
obscure psychophysiological relation in terms of the practically ascertainable
relations of causality and expression. But it seems more impor-tant from the
point of view of the causal theory of psychophysical time to consider the
statement in question as directly defining the expressive relation in terms of the
other two since in this way we bring out the causal character of the expressive
relation and of the method of psychophysical synchroniza-tion which it employs.
In particular, we derive from this interpretation that every psychological state that
has a bodily expression also has a physiological substratum, proving that every
case in which the synchronization is established by means of the expressive
relation can also be treated by employing the psychophysiological relation, and
that, consequently, the definition of psychophysical time based on this last
relation is generally valid.
It will be noticed that neither of the two practical methods of psycho-physical
synchronization, taken by itself, allows us to localize a psychological state very
precisely in psychophysical time (this would be possible if we could establish
which position of the hands of a sufficiently precise clock is simul-taneous, in the
psychophysical sense, with this psychological state): the perceptual method is
applicable only to physical events which are prior to a given psychological state,
and the expressive method serves only to establish the posteriority of a physical
event with respect to this state. It is only by
246 DURATION AND CAUSALITY

the joint use of the two methods that we can establish that a psychological state
which is later than a physical event El (this is established with the first method) is
earlier than physical event E2 (using the second method), thereby enclosing the
psychological state in the physical interval E 1 E 2. 158 By trying to make this
interval smaller and smaller, we will locate the psychological state in
psychophysical time more and more precisely. At the limit, whi~h is impossible
to attain, the two ends of the interval E 1E 2 , (i.e., the physical event perceived
prior to the psychological state and the physical event ex-pressing this state and
succeeding it in time) would coincide and become simultaneous with the
physiological substratum (of the psychological state) and thus the localization in
time would become absolutely rigorous. [This follows from the well-known
phenomenon of chronaxie i.e. the time a neural current of twice the threshold
stimulus strength must flow to excite tissue response. 159 ] We see that the
psychophysiological method of synchroniza-tion, the only theoretically
satisfactory one, since it provides precise and not approximate data, can be
conceived of as the common limit approached by the other methods, as the
precision with which they localize psychological events in time increases.

To illustrate the significance of these different methods of psychophy-sical


synchronization and the difficulties inherent in their application, we shall
examine another typical example, that of dreams. Is it possible to localize them in
time? The perceptual method is of little help here, since the dreamer himself
would have to apply it. It might aid in localizing the dream as a whole, if, for
example, the sleeper remembered having heard the clock strike a certain hour at
the beginning of the dream or upon awakening. But such a summary localization
is obviously insufficient. We do not believe that the dream as a whole floated in
time between the limits of the interval occupied by sleep, and we require that
each phase of the dream which took place during this interval should be
localizable at a precise instant of the interval. We would therefore like to be able
to answer the following two questions: (1) at what moment did a particular phase
of the dream (which the dreamer remembers after awakening, such as a sudden
fright) take place? (2) what phase of the dream took place at a given instant of
the interval? Admittedly, these two questions would ordinarily be considered
equivalent, i.e., a complete answer to one would also clarify the other, since we
do not believe that temporal dissociations of the kind described in Section I occur
during sleep.
In certain cases the sleeper will easily localize in time a particular phase of his
dream, a sudden fright, for example. If he had had some perceptions, even
disfigured by sleep (e.g., if he remembered having heard a bell while
NON-PHYSICAL TIME 247

experiencing this fright), the perceptual method would be applicable here.


The same would be true if he had been conscious of visceral sensations,
since, for example, by counting his heart-beats from the time he went to
sleep he might be able to localize temporally the phase in question. But all
this is highly improbable. In general, the perceptual method will be
inapplicable during the period of sleep, and the sleeper, once awake, will be
able to remember the unfolding of his dream, but no amount of introspection
will enable him to localize its phases in time.
There remains the expressive method. The sleeper can be observed while
he sleeps, and the expression on his face at the moment when he experiences
a fright would reveal his state of mind to the observer; this state of mind
could also be expressed by an involuntary exclamation. This expressive
method is also applicable independently of any external observer, since he
could be replaced with a machine which would automatically register the
bodily phenomena accompanying the fright of the sleeper and thus conserve
their trace, which th~ sleeper, once awake, could use to localize his dream in
time. If these bodily phenomena were lacking, the problem would become
practically insoluble.
It might be said that the problem would then also lose its theoretical meaning:
in saying that the psychological states of the dreamer find no bodily expression,
we thereby assert that no one besides the dreamer himself can know anything
about his psychological life, since our knowledge of the psychological life of
others rests precisely upon the expressive relation. Furthermore, since, by ruling
out the perceptual method we render the introspections of the sleeper useless, it
follows that the temporal localization of the dream is impossible, since this dream
is supposedly inaccessible to the dreamer and to those who observe him. This
objection is not conclusive, however. We have assumed only that the
psychological states of the sleeper fInd bodily expression neither during nor after
sleep. Once awake, the sleeper will be able to communicate his dream to the
person who would have observed him while he slept, and then they will both be
incapable of assigning a place in time to this dream by means of the joint
perceptual and expressive methods.
Undoubtedly this diffIculty will seem practically insurmountable to them.
But, theoretically, we can still attempt the psychophysiological method.
According to the hypothesis of psychophysical parallelism, someone who
observed the brain of the sleeper while he slept and who knew the one-to-one
correspondence between the psychological states and their physiological
substrata could infallibly deduce the corresponding psychological states from
the successive states of the brain. Thus formulated, this hypothesis is
248 DURATION AND CAUSALITY

undoubtedly subject to criticism. But we shall see later that to synchronize the
universe of the mind with that of the body, it suffices to accept a restricted
parallelism, according to which only certain psychological states can be partially
reconstructed, starting with cerebral data. Relying on this restricted parallelism
and taking certain precautions, we could still localize the dream in time, if we
assumed that at each instant of the dream there was at least one psychological
element having a substratum. 160 If this hypothesis is not verified, if, for example,
the fright in question is not simultaneous with any psychological phenomenon
having a physiological substratum, the third method of synchronization will also
fail. I believe that, in this case, the problem is practically and theoretically
insoluble, and, that further reflection will show that it is badly stated since no
sense can be attached to the temporal localization of the psychological state in
question.

3. PSYCHOLOGICAL AND INTER-PSYCHOLOGICAL TIME

We shall now consider psychological and inter-psychological time insofar as


they differ from intuitive time. Any succession or simultaneity between
psychological states belonging to the same self is psychological, by definition
- but it is intuitive only if it corresponds to a perceived temporal figure or to a
particular chain of recollections. However, going back to sections (2) and
(3) of Chapter VIII, we see that the psychological temporal order is in general
defmed by the intuitive order, save in exceptional cases when the latter contains
gaps of the second kind. The same is not true of inter-psychological time.
Undoubtedly, the inter-psychological time or order of phenomena can be
considered as intuitive insofar as it can be identified with the intuitive order of
the perceptions of these phenomena (Cf. Chapter VIII, Section 5). But it is only
an exception that two states belonging to different fluxes of consciousness are
'perceived' by the same observer. Interpsychological simul-taneity and succession
in general will not be intuitive, and the problem is that of how to delimit the
sense of before, after and during, so that it will no longer be restricted to
intuitively knowable cases and will be applicable to any pair of psychological
states.
The answer is contained in the definition of the psychophysical temporal
relations. In comparing a pair of psychological states with suitably chosen
physical events, we will deduce the psychological or interpsychological relations
characterizing this pair from the psychophysical temporal relations linking it with
physical events. Thus, two states which are psychologically simultaneous with
the same physical event will be considered (psychologically
NON-PHYSICAL TIME 249

or inter-psychologically) simultaneous with each other. We can formulate other


rules which allow us to reduce the temporal order of psychological states to the
psychophysical order. We will adopt here a simple formulation which uses
explicitly neither psychophysical succession nor simultaneity but is based
directly on the concept of a physiological substratum: psychological state X is
simultaneous with (or earlier than) state Y, if the substratum of X is simultaneous
with (or earlier than) the substratum of Y.
This defmition, which brings out the fact that it is through its connection with
the material universe that psychological reality acquires a temporal structure,
raises many problems. In the first place, one may ask if it contains the defmition
of intuitive succession and simultaneity as a particular case, or, in other words, if
the intuitive succession and simultaneity of two psycho-logical states belonging
to the same self are also definable in terms of the succession or simultaneity of
their physiological substrata. This would amount to agreeing with Spinoza's
thesis that the order of psychological facts is only a copy of the order of
physiological processes. But science admits of several relations between these
orders. For example, if physical event A precedes (in the psychophysical sense)
psychological state B, and if B (intuitively) precedes state C, then even A
(psychophysically) precedes state C. This statement is not a tautology, but an
empirical generalization. But can we conclude from it that the succession of two
psychological facts is equivalent to the succession of their physiological
substrata? In view of the relativity of physical succession, this seems doubtful to
us. We know that in general two events which succeed each other with respect to
a given reference system no longer do so with respect to some other system.
Therefore, it seems difficult to say that the succession of two physical events,
which is defined only with respect to a given system of reference, can be
equivalent to the succession of two psychological events, which implies no
dependence on any system of coordinates. We might try to overcome this
difficulty in two ways. First of all, we may try to single out a privileged
coordinate system, to which the psychological life of a given individual would be
more closely linked, saying that the order of psychological states corresponds to
the order of their physiological substrata when related to this privileged frame of
reference, (we will, of course, have in mind a frame of reference attached to the
individual's body). Notice, however, that such a system would be completely
indeter-minate, since the parts of the body move with respect to one another and
there are always several motions going on inside it. It is therefore impossible to
rigidly attach a frame of reference to the entire body. The possibility of
constructing an intermediate time, common to all parts of the body, seems
250 DURATION AND CAUSALITY

questionable. Secondly, we might try to attach the psychological temporal order


to an invariant physical order, making use of the fact that two events which
succeed each other in one coordinate system and which are causally related, also
succeed each other in every other admissible system. The exist-ence of a chain of
recollections between two psychological states would, from this point of view,
correspond to the existence of a causal relation between their substrata. However,
this hypothesis would not explain the correlation between intuitive and physical
simultaneity. We know that physical simul-taneity, which is relative in principle,
becomes invariant only for events which coincide in space. But the cerebral
localization of psychological functions seems to suggest that the physiological
substrata of two psychological states are in general located at differeent points of
the cerebral cortex; their simul-taneity is therefore always relative; and from this
we conclude that the invariant simultaneity of psychological states cannot be
equivalent to the relative simultaneity of their physiological substrata.

It seems, therefore, that a rigorous parallelism of the temporal order of


psychological states with that of their physiological substrata does not exist. We
shall not examine what consequence this has for the general problem of
psychophysiological parallelism. We shall say only that the difficulties we have
just raised could be removed if, instead of a rigorous temporal corre-spondence
between the two orders of phenomena, linking every instantaneous phase of a
psychological process with a determinate phase of its substratum, we assume an
approximate correspondence, according to which psychological states extended
in time would correspond to physiological processes taken as wholes, without
this global correspondence involving a one-to-one corre-spondence between the
consecutive phases of the two series. 161 It is precisely the imprecision of this
correspondence which would allow us to correct for the discrepancies resulting
from the relativity of physical time. Admittedly, these correlations should not go
beyond certain limits, which is what would happen if we considered every
reference system admissible. This system is therefore not arbitrary, and it seems
probable that it does not differ greatly from the system attached to the Earth,
since it is by using such systems that we came to establish the approximate
correspondence between intuitive and physical time.

The employment of the psychophysiological relation in the theory of


psychological time therefore falls quite short of removing all the difficulties.
Nevertheless, its use seems to be unavoidable. To show this, let us try, ideally, to
eliminate from the organism associated with an individual flux of con-sciousness
all that can be removed without interfering with the flux itself. If
NON-PHYSICAL TIME 251

it is true that psychological states depend immediately only on what takes place
in the brain, then bodily phenomena occurring outside the brain can influence the
psychological states associated with the organism only by influencing the regions
of the brain which correspond to these psychological states. It follows from this
that the psychological life of an individual would not change if we removed the
brain, provided that its boundary conditions were kept the same as they had
normally been as a result of the brain's relations with the other neighboring cells
of the organism. It is clear that the only method available for synchronizing the
flux of consciousness associated with the singularly reduced organism, which is
incapable of communicating with us by means of expressive motions and of
participating in his physical environment by means of the perceptual relation, is
the psychophysiological method. Only a knowledge of the psychophysiological
relation would enable us to discover and locate in time the psychological states
associated with the states of the isolated, brain. [The above consideration
obviously neglects the relevance of a human organism's endocrine condition to
the temporal organization of his mental experiences. H. M.]

4. UNIVERSAL TIME

We have seen in the preceding sections that the temporal concepts of succes-sion
and simultaneity differ in meaning depending on whether they apply to a pair of
physical events, or to a pair of psychological states (whether or not these states
belong to the same self), or finally to a mixed pair. It follows that science uses
four distinct temporal orders, which we have designated as physical time,
psychological time, inter-psychological time, and psycho-physical time. Yet,
history, geology, astrophysics, and paleontology seem to locate all events,
whether physical or psychological, in a single universal time. How does this
come about? I believe that this time unique to science is neither superimposed nor
juxtaposed to the physical, psychological, inter-psychological, and
psychophysical times, but that it is quite simply their synthesis. More precisely,
event A precedes event B in the sense of universal priority if it precedes it in the
physical, the psychological, the inter-psycho-logical, or the psychophysical sense.
And, similarly, universal simultaneity must be def'lned as a 'relational sum' of the
four particular simultaneities.
The question now is how the properties of universal time can be deduced
from the properties of its components, or, to put it differently, what are the
properties of these components that they result in a continuous and unidimen-
sional order of instants. We shall take physical time as our point of departure,
252 DURATION AND CAUSALITY

taken by itself, it already forms a unidimensional continuum of instants, and


we will try to show that in order that physical time not change its structure
with the addition of the extra-physical components, it is necessary and suf-
ficient that the principle of psychophysical parallelism be valid.
From the relational point of view, we must take account of the fact that,
since the instant is defmed as the set of events simultaneous with a given
event, the necessary and sufficient condition for the unidimensionality and
continuity of the order of instants is that simultaneity, which is supposedly
symmetrical, transitive, and incompatible with succession, conceived as an
asymmetrical, transitive, and continuous relation, together should form a
connected whole, that is, of any two events, one is always earlier than, simul-
taneous with, or later than the other. Translated into relational language, the
problem of the deduction of the properties of universal time from those of its
physical and extra-physical components can be formulated in the following
manner: if we assume that physical simultaneity and succession have the
above-mentioned properties, guaranteeing the linear character of the physical
temporal order, how can we be sure that universal succession and
simultaneity, each defmed as a function of its particular components, will
have the same ordinal properties?
In its ordinary form, the principle of psychophysical parallelism implies
the following three statements: (0:) every psychological state has a physio-
logical substratum, ({3) if two psychological states succeed each other (in
the intuitive sense), their substrata succeed each other (in the physical sense),
(y) if two psychological states are simultaneous (in the intuitive sense), their
substrata are simultaneous (in t.he physical sense). Referring back to our
defmitions of inter-psychological time, it will be noticed that these three
statements can be replaced with a single proposition, asserting that intuitive
simultaneity and succession are, respectively, special cases of inter-psycho-
logical simultaneity and succession, which very nicely brings out the con-nection
between the structure of universal time and the principle of psycho-physical
parallelism. With the aid of these statements, we can simplify the definitions of
universal succession and simultaneity by putting them into the symmetrical form
of inter-psychological succession. It will be remembered that the inter-
psychological succession of two psychological states was defined as the
succession of their substrata. Provided that we give a suitable generalization of
the concept of substratum, we can defme universal succes-sion in exactly the
same manner, that is, by calling the generalized substratum of a phenomenon the
phenomenon itself, if it is a physical phenomenon, and its physiological
substratum, in the ordinary sense, if it is a psychological
NON-PHYSICAL TIME 253

phenomenon. It is easily verified that by virtue of statements a, {3, and 1, the


definition of universal succession given above is equivalent to the following
definition:
Phenomenon X is earlier than or simultaneous with phenomenon Y (in
universal tiine), if the generalized substratum of X is earlier than or simul-
taneous with the generalized substratum of Y (in physical time). We see
immediately that, thus defined, universal succession and simultaneity have
all the properties of physical succession and simultaneity which assure
physical time its structure as a unidimensional continuum and that,
consequently, universal time possesses the same structure.
This derivation of the structure of universal time is not very satisfactory,
however, because the principle of psychophysical parallelism upon which it is
based is open to criticism in several respects. The statement that every
psychological state has a physiological substratum, often maintained for more or
less metaphysical reasons, [the principle is probably traceable to Leibniz'
metaphysical view of a preestablished harmony between the monads termed
'mind of X' and 'body of X', respectively. H. M.J seems to be in no way verified,
or even suggested by experience. One might certainly waver on the question of
whether a particular class of psychological states do or do not have physiological
substrata - whether, for example, Wundt's statement that only simple sensations
and feelings have substrata 162 is correct. The con-siderations advanced in
Chapter III, Section 2 prove, in any case, that only psychological states fmding
bodily expressions have substrata. But it seems very doubtful that all
psychological states, even the most complex and the most devoid of sensory
elements, should also have physiological substrata. It is therefore desirable to
deduce the structure of universal time from a state-ment which is more restricted
than a, assuming that all psychological states do not have substrata. This is in fact
possible, as is suggested by the following table, in which the italic letters in the
first two columns represent the psycho-logical states belonging to two different
fluxes of consciousness respectively, while the Greek letters in the third column
stand for the physiological sub-strata attached to the states symbolized by the
corresponding italic letters. The occurrence of two letters on the same line, or on
different lines, indicates the simultaneity and succession respectively of the
corresponding phenomena:
abc d e

a b' d' e' a' {3' l' {)' e'


'

I II III
254 DURATION AND CAUSALITY

Thus, psychological state a is intuitively simultaneous with states band c, inter-


psychologically simultaneous with d, and psychophysically simultaneous with its
substratum, lX, and also with events {3, r, O. Event lX', physically simul-taneous
with {3', is psychophysically simultaneous with state a', of which it is the
substratum, and psychophysically earlier than state a. Assuming the
psychological and physical temporal orders to be given, we admit that the
arrangement of letters within each column is fixed, and, taking physical time as
our point of departure, we assume that the third column cannot be dis-placed
vertically as a whole. But these two assumptions determine neither
psychophysical time nor inter-psychological time, since they do not fix the
position of the first two columns with respect to the third: by lowering the first
column so that the letters a, b, c, are brought into the same line as d' and e', we
change the inter-psychological simultaneity of a with d and the psychophysical
simultaneity of a with lX, into succession. Therefore, deter-mining the structure of
universal time amounts to moving the three columns vertically so that the letters
occurring on the same line always represent simultaneous phenomena. The
diagram shows that, for this purpose, it is sufficient to determine psychophysical
time, since, when this is done, the inter-psychological temporal order is
determined (by fixing the positions of the first and second columns with respect
to the third, we also fix their reciprocal position with respect to each other.) We
also see that it is not at all necessary for all psychologically simultaneous states
a, b, c to have substrata lX, {3, r: if only one of them has a substratum, the location
of all of them will be fixed, since, by virtue of their intuitive simultaneity, the
others will all be on the same line. Therefore, the Table suggests that, in order to
guarantee the linear structure of universal time, it will suffice to replace statement
lX, which requires every psychological state to have a physiological substratum,
with the more restricted statement lX', according to which every set of intuitively
simultaneous phenomena contains at least one phenomenon having a sub-stratum,
or, in other words, in each flux of consciousness, there is, at each instant, at least
one phenomenon having a substratum. 163 In order to apply this postulate, it is
evidently necessary to assume that intuitive simultaneity is transitive, and to
change the defmitions of extra-physical time, substituting the following
statements for those above:

Psychological state X is psychophysically simultaneous with physical event Y


if it is intuitively simultaneous with a state whose substratum is physically
simultaneous with Y.
Psychological state X is psychologically or inter-psychologically simul-
taneous with psychological state Y if it is intuitively simultaneous with a
NON-PHYSICAL TIME 255

psychological state whose substratum is physically simultaneous with the


substratum of a state intuitively simultaneous with Y.
By substituting these definitions of extra-physical simultaneity in the
definition of universal simultaneity and by similarly replacing in the definition of
universal succession the concepts of extra-physical succession with concepts
analogous to those just defined, we obtain a definition of universal time which is
a bit complicated, but which nevertheless has the virtue of agreeing quite closely
with the facts furnished by experience. We shall see that in fact, when related to
these definitions, the restricted principle of psychophysical parallelism is
sufficient for the derivation of the linear structure of universal time from the
properties of physical and psychological time.
To demonstrate asymmetry we need only remark that since universal suc-
cession is defmed as a sum of four relations, its asymmetry will be guaranteed if
each of these four relations is asymmetrical. The first is so by hypothesis, and the
asymmetry of the other three follows from the restricted principle of parallelism.
To show the asymmetry of psychophysical succession, for example, let us assume
that psychological state X succeeds physical event Y, i.e., that the substratum of a
state intuitively simultaneous with X physically succeeds Y. If Y also succeeded
X, it would have to succeed physically the substratum of a state X" intuitively
simultaneous with X. States XX", both simultaneous with X, are intuitively
simultaneous with each other, by virtue of the transitivity of intuitive
simultaneity, and their simultaneity entails that of their substrata, by statement (3.
Y, although earlier than one of these substrata, would then succeed the other, in
spite of the fact that these two are simultaneous with each other. This is
impossible, since it follows from the hypotheses about physical time that an
event ~ earlier than 1/, is also earlier than ~, if 1/ and ~ are simultaneous. 164 By
using statement -y, we can show in the same way that psychological (and inter-
psychological) succession is asymmetrical.

The transitivity of universal succession means that every phenomenon X,


which succeeds a phenomenon Y which in turn succeeds phenomenon Z, also
succeeds Z, the variables 'X', 'Y', and 'Z' designating physical events, or states
belonging to a single conscious flux, or to different fluxes. We can distinguish
eight possible cases by noting which of the three variables designate a phy-sical
event. If two consecutive variables designate psychological states, we will then
have also to distinguish those cases in which these states belong to the same flux
or to two different fluxes. We thus obtain thirteen different cases, according as
'X', 'Y', 'Z' deSignate three physical events (which reduces universal succession
to physical succeSSion), or 'Z', a psychological state, and
256 DURATION AND CAUSALITY

'X' and' Y', physical events, etc .. In each of these thirteen cases, statements a, {3,
'Y entail that X is earlier than Z. Just to fIx our ideas, let us consider the case
where X is a physical event, while Y and Z designate psychological events
belonging to different fluxes. The transitivity of universal succession here
amounts to saying that if event X psychophysically precedes state Y, and Y
interpsychologically precedes Z, then X psychophysically precedes Z. This is
because, according to the definition on p. 254, X physically precedes the
substratum S of a state intuitively simultaneous with Y, and the substratum S' of
another state intuitively simultaneous with Y physically precedes the substratum
S" of a state intuitively simultaneous with Z. But Sand S' corre-spond to states
intuitively simultaneous with a third state, and are therefore also simultaneous
(with each other). Since X occurs before S, and S' before S", X occurs before S",
which means that S psychophysically precedes Z. In the same way, we can verify
that in the other cases the restricted principle of parallelism entails the transitivity
of universal succession.
We shall demonstrate the transitivity of universal simultaneity in precisely
the same way. Its symmetry follows from the symmetry of the four particular
relations of which it is the sum. The incompatibility of universal succession
with universal simultaneity results from the fact that each component of the
latter excludes the homonymous component of the former, by virtue of the
properties of the particular time to which they belong, as well as the other
three components which, because of the diversity of its members, can be
linked by different components. Finally, the continuity of physically succes-
sion directly entails that of universal succession. We thus see that the
restricted principle of psychophysical parallelism, along with suitably
modifIed defIni-tions of extraphysical time, is suffIcient to prove the linear
character of universal time.
What is the importance of such a derivation? The hypothesis of psycho-
physical parallelism, even in the restricted form which we have employed to
deduce the properties of universal time, is far from being unquestionable.
There is no doubt that the arguments drawn from causal continuity and the
physiological dependence of sensations render it plausible, but here, as every-
where, it is possible to replace the given hypothesis with another without
contradicting the facts which verify it. Perhaps it will be said that to base the
temporal order on a probable, but, after all, uncertain hypothesis is to give it a
very shaky foundation, and that the fundamental laws of time, concerning the
properties of succession and simultaneity, seem to enjoy, so to speak, an
epistemological dignity far superior to that of the parallelistic hypothesis
upon which we make them depend. However, the analysis of the
NON-PHYSICAL TIME 257

extra-physical temporal order, especially of its psychophysical and inter-


psychological components, clearly shows that what we really know about
this order rests upon the psychophysiological relation, that the only methods
employed by science which effectively permit us to locate psychological
states in universal time are contained in the psychophysiological method. If
the hypotheses upon which this method rests are uncertain, so is our knowl-
edge of the localization of phenomena in universal time and of its funda-
mental laws. We have no a priori evidence that these laws must be modelled
on the laws of the time and revealed by perception and memory. No one
would maintain that restricted psychophysical parallelism is true a priori;
neither, therefore, should we affirm a priori that universal time exists, since
under certain conditions the hypotheses concerning its fundamental
properties are equivalent to the principle of psychophysical parallelism.
Moreover, this general conclusion is imposed by an examination of the
facts themselves, independently of our derivation. The example of the
sleeper, analyzed above, shows that we have no a priori certainty that it is
possible to localize dreams precisely in time, and that the only argument
capable of lending credibility to the ideal possibility of such a localization
rests upon psychophysiological considerations. 165 The law which requires
that every phenomenon contained within an interval of time (Le., later than
its beginning and prior to its end) be localizable in a precise manner within
that interval, whether its duration extends over a particular part of the
interval or over all of it, is a direct consequence of the fundamental
properties of time. However, we have no argument beyond that drawn from
parallelism, to rule out the view that a psychological state may, so to speak,
float in an interval of time without occupying a fIxed place in it.
Moreover, as with any empirical hypothesis, the principle of parallelism
only replaces a vaguer approximation with a more precise one, without ever
attaining absolute precision. We have seen that the relativity of physical time
precludes an exact correspondence between a physiological series and a
psychological series, Le., a correspondence which would link each
instantane-ous phase of the latter with a unique instantaneous phase of the
former. A global correspondence, linking processes extended in time and
allowing us to correct for the relativity, will lead only to an approximate
localization of the psychological state, but this approximation will be
essentially better than that obtained by methods which do not directly use the
physiological substratum of the state in question. 166
[An important role in the merger of the four distinct temporal orders has been
attributed to the principle of psychophysical parallelism. Accordingly,
258 DURATION AND CAUSALITY

it may be worthwhile to mention a few more or less recent developments


related to this principle. In his treatment of quantum mechanics, J. von
Neumann showed in 1932 that this theory is compatible with the principle. In
addition, three different approaches to the principle have been success-fully
made in the last decades: (1) Penfield had appreciably extended the scope of
the principle, on the basis of his studies of epileptic cases made during the
Second World War. More specifically, he succeeded in specifying the sub-
cortical center for the memory of words. (2) Rechtschaffen and associates
have extended the cerebral substratum for mental experiences occurring
during sleep. (3) Ingle and his followers have investigated the psychological
effects of glands of internal secretion. All these advances have appreciably
strengthened the observational basis of psychophysical parallelism. Needless
to say, the aforementioned investigators were not in a position to determine
the true nature of the relation between a psycho-
logical experience and its physiological substratum. The relation might
consist in numerical identity defined by Russell in terms of membership
in the same classes. But a more subtle concept of 'synthetic identity', in-
troduced by H. Feigl in this impenetrable, theoretical area, may be more
adequate. However, for the substantiation ofthe merger of the four temporal
orders, no assumption about the nature of the psychophysiological relation is
needed. Psychophysical parallelism, construed as a many-one relation
between the physiological processes in a human organism and the psycho-
logical experiences associated with this organism, perfectly suffces for our
purpose.
In view of the particular importance of the principle of psychophysical
parallelism, it may be helpful briefly to summarize the fmdings of A. Recht-
schaffen, condensed in an essay entitled 'Current Research on Sleep and
Dreams' (1965). Two types of sleep are distinguished (1) the sleep accom-
panied by a rapid eye movement of the person asleep (REM); (2) sleeps
devoid of this physiological accompaniment. A basic tool in this area is
obviously the electroencephalograph. The chemistry of sleep is reasonably
established and explicitly described in A. Rechtschaffen's monograph. The
dream state of REM is lucidly and competently described. So is the develop-
ment of infant sleep. Control over dreams and other states of consciousness is
specified. This chapter of the monograph includes sections devoted to sleep
learning, to discrimination during sleep, performance during sleep,
habituation, sleep habits, the awakening process etc. There is a fascinating
chapter entitled 'The Meaning of Dreams', a sub-section of which deals with
influencing dream-content, the dream-recall, the intensity of dream and the
NON-PHYSICAL TIME 259

stimulation before sleep, a classification of dreams and somewhat understated


'speculations about the neural basis of dreams'.
The aforementioned titles of sub-sections can give a good idea of the present
reach of dream research. The last subtitle characterizes the epistemo-logical level
of the essay: while dealing with dreams, the essay shows a rare, epistemological
carefulness in the statements it makes, a mastery of con-temporary literature and
of the use of appropriate tools. It is difficult to study the essay closely without
reaching the conclusion that a most significant plunge has been made into the
depths of psychological life, and, indirectly, of psychological time.

Oscar Wilde once stated that the best way of resisting temptation is yield-ing.
Taking advantage of the wisdom of this British writer, I should like to return to
Penfield in order to add some illuminating quotations from his above-mentioned
studies. The Montreal scientist draws a line between sensory and motor
responses on the one hand, and 'psychological' responses on the other.
Hallucinations and recollections belong in the latter class. They are

at times obtained by electrical stimulation of the cortex on the superior and lateral surfaces
of the temporal lobe of either side. In general, the recollections produced by electrical
stimulation seem to be as clear as they would be a few seconds after the experience. In fact
they are apparently as clear as they were during the experience ...
after ten years of absence the subject may be quite unable to recall a former friend clearly.
Nevertheless when the latter is met again, however unexpectedly, one recognizes him and
sees at once changes in minute details ... This new experience in which the friend was the
centre of attention was immediately compared with the ganglionic records of that friend
which had been ostenSIbly forgotten but not lost.

It is profitable to compare these lines of W. Penfield with what Henri Bergson


once wrote in Matiere et memoire. He also maintained that the memory of an
actual experience is never lost. But he firmly denied the cerebral localization of
this memory.]

5. CONCLUSION

This approximate character of the knowledge of universal time merely serves to


confrrm its causal character, which emerges from the whole of the present
study. We have seen that the fact that the four components of the temporal order
of phenomena, (which we have designated as physical, psychological,
psychophysical, and inter-psychological time) define a unique, continuous and
unidimensional temporal order, is equivalent to a particular property of the causal
interconnection of phenomena, and that each of these components
260 DURATION AND CAUSALITY

rests in tum upon causal structures. This is true of psychophysical time, which is
knowable by the three methods whose causal character we have pointed out, and
also of psychological and inter-psychological time, a knowledge of which can be
derived from a knowledge of psychophysical time. The funda-mental properties
of physical time have likewise been deduced from those of the causal relation.
Coincidence, simultaneity, succession, and duration have all been defined in our
axiomatic system by means of the method of the causal decompositions of
becoming. This method, based on symmetrical causality (which is definable
independently of time and indifferently links cause to effect and effect to cause),
illustrates the isotropic character of physical time, which is connected with the
reversibility of the causal laws of physics. It explains qualitatively and
quantitatively the relativity of physical time by showing that this relativity
consists in the multiplicity of causal decompositions of becoming, and that the
Lorentz formulas which mathe-matically express it can be deduced from some
hypotheses on the causal relation. The employment of light signals and rigid
systems of reference, which usually serve to define temporal relations and
demonstrate their relativity, is found to be a practical procedure with restricted
applicability and in no way inherent in the temporal structure of becoming.

This structure is determined by the causal interconnection of phenomena,


independently of all temporal intuition, that is of the intuitive sense of
before, after, and during. We have seen, however, that intuitive time, itself
restricted to fragments of psychic becoming and profoundly different in its
structure from universal time, plays an indispensable role in the knowledge of
universal time: the causal temporal order of phenomena would be hardly
knowable if in certain cases it were not conventionally identified with the
intuitive temporal order of their perceptions.
SUPPLEMENT

1. THE PRESENT EMPIRICAL STATUS OF PSYCHOPHYSICAL


PARALLELISM

I shall wind up my analysis of the merger of four types of temporal order with an
exploration of the principle of psychophysical parallelism underlying this merger.
The evolution of the mind-body problem shows clearly that a considerable part of the
relevant controversies is traceable to the vagueness of the terms used by the
formulation of the issue. More particularly, this holds true of the view of the mind-
body relation which became almost dominant at the beginning of our century and has
been termed 'psychophysical parallel-ism' .167 My comments in the preceding section
have aimed at a precise formulation of the three basic postulates inherent in
psychophysical parallel-ism and at a determination of their mutual relationships. This
section deals with the current status of psychophysical parallelism.

Psychophysical parallelism originated with the advance of empirical science.


(Like most major scientific advances, the principle of the psychophysical is
nevertheless traceable to a remote metaphysical doctrine. I have in mind Leibniz's
view that a mind animating a particular organism forms with the latter a pair of
monads which are both governed by the principle of 'preestab-lished harmony'.)
Psychophysical parallelism was not a revival of ancient controversies related to
the mind-body problem and supported by arguments derived from a pre-scientific
meaning-analysis of the relevant terms and pre-scientific, observational data
concerning the mind-body problem. Rather, it was an attempt to generalize a set
of well-established scientific laws relevant to the mind-body issue. There are
many scientific laws that deal with neither psychological events nor phYSical
events exclusively. Such laws are exemplified by the Fechner-Weber rule which
associates changes of the intensity of the physical stimulus with changes of the
corresponding sensation. They do not belong either in psychology, or in
physiology proper. Similar laws govern the origin of various perceptual
delusions, e.g., optical delusions. It is impossible to formulate these laws either in
purely physical or in purely psychological terms. This also holds for laws which
associate specifiable qualities of a type of sensory experiences with
corresponding properties of the physical stimuli
261
262 SUPPLEMENT

(for instance, the pitch, timbre, and volume of an auditory sensation with
length, shape and amplitude of the corresponding physical vibration). Less
familiar examples are provided by the 'temporal lobe syndrome' described by
H. Kluver, in 'Mechanisms of Hallucinations' and 'Neurophysiology of
Perception' .168
In addition to a conceptual analysis of the principle of psychophysical
parallelism, it is useful to recall the basic facts established so far by
investiga-tions of the cereblal cortex of man, of the physiological substrata of
human dreams and of the psychological effects of glands of internal
secretion. These three lines of approach can be associated, respectively, with
three groups of scientists, to be referred to by listing three groups of
representative names: (1) J. C. Eccles, W. Penfield, W. R. Sperry; (2) N.
Kleitman, A. Rechtschaffen, W. B. Webb; (3) D. Ingle. 169
Penfield has apparently obtained his impressive amount of information
about the physiological substrata of mental experiences from the thousands
of cases of epilepsy whose study was made possible by the Second World
War. His basic finding may be put as follows: there is a sub-cortical part of
the brain which produces the fundamental phenomena of human
consciousness, voluntary activity, memory of past experiences, intellectual
activity and planned action. This part of the brain, referred to by Penfield as
the 'cen-terencephalic system', consists of all the brain except the cerebellum,
the cerebral cortex and their dependencies. By definition, those parts of the
brain just described, and often referred to as the 'brain stem', do not include
the elements of the brain stem which have no symmetrical connections with
both cerebral hemispheres.
Reasonably accurate statements are then made by Penfield concerning the
various sensory representations in the brain (auditory, somatic, visual), the
cerebral motor apparatus including the set of voluntary motions, the cortical
mechanisms of recollections and hallucinations (localized on the superior
and lateral surfaces of the temporal lobe of either side), the distinction
between the functions of recording perceptions and of storing these records,
etc. The philosophically most significant aspect of Penfield's fmdings is not
related to sensory and motor experiences since both were known to have
cerebral substrata and were admitted in this capacity by philosophers whose
contri-bution to the mind-body problem is most highly regarded (for
instance, H. Bergson's work Matiere et memoire based to a considerable
extent on the classification by P. Marie of his seven types of aphasia). Some
of Bergson's claims have actually been confirmed by Penfield. This applies
e.g., to the assumption that man never forgets what he has ever perceived
although he
SUPPLEMENT 263

may be incapable of arousing the relevant recollections. In contradistinction


to Bergson, however, Penfield has succeeded in establishing cerebral
substrata for mental experiences which are neither of the sensory nor of the
motor variety. More specifically, the discovery of the cerebral substrata of
memories is at variance with Bergson's entire outlook and strengthens
considerably the principle of psychophysical parallelism.
The findings of W. R. Sperry are somewhat less comprehensive, but in a
sense, more dramatic. He succeeded in shedding new light on the psycho-
logical significance of the two cerebral hemispheres, and the 'corpus
callosum' which connects them in normal cases, with no surgical
interference. Sperry's discoveries deal both with humans and some sub-
human animals. The dramatic aspects of his findings become manifest in
those cases where the connection between the two cerebral hemispheres is
impaired, either because of a surgical intervention, or for other, biological
reasons. In these cases, not only single, mental events in the life of a human
individual are affected, but those deeper layers of psychological life, which
can only be understood in terms of a personality-structure, are also involved.
I am not referring here to the Freudian multiplicity of selfhoods, which tends
to replace the single, human personality with a trinity of sorts. What Sperry's
exploration of bisected brains shows is that the 'dominant' left hemisphere,
where speech is centered, may be duplicated when the connecting corpus
callosum is impaired. It may still be an open question whether the result of this
impairment is a split personality or a shift of the cerebral personality-substratum
from one hemisphere to the other one. It is not quite clear, either, how the two
hemi-spheres manage to support a single personality in all normal cases and what
light if any is shed on schizophrenics and related, abnormal humans.
But the important point is that the neurophysiological basis of man's
mental life affects not only the conscious episodes occurring in his wakeful-
ness and his dreams. The more pervasive and enduring features of mental life
which determine the uniqueness of a personality and its possible splits, are
also brain-dependent. The cerebral localization of sensory and motor experi-
ences is demonstrated in a most illuminating way when (regrettably) an
experimental animal is categorized as a 'sensory nerve cross'. In these situa-
tion, a surgical cross-union of the sciatic nerve brings about a switch in the
nerve connections from a (hind) left leg to a (hind) right leg. The animal then
reacts to imputs in one leg in the way it would have reacted to a similar input
in the other leg, should the sciatic nerve not have been re-oriented. Actually,
not only literally sensory and motor events are involved. The changes in
localizability apply also to experienced pain and pleasure.
264 SUPPLEMENT

It is certainly implausible to deny the causal interaction of mental and


neurophysiological events in situations described by Sperry. Strangely enough,
neither he, nor the other scientists who have contributed to the 1976 volume,
mention the relevance of the quantal surrender of physical determinism to the
status of the interactionist hypothesis. And it is not very helpful that (in
contradistinction to Eccles) Sperry formulates the conclusion derivable from his
findings as a 'holistic' 'emergentism'. I have not discussed the issue of
emergentism of any empirical theory with regard to some other theory, e.g., of
psychology to neurophysiology, or, for that matter, of phenomenological
thermodynamics to the combined theories of Newtonian mechanics and
Maxwellian electrodynamics, simply because the essential role of operational
definitions or defmitional criteria entails classical emergentism. If a theory T
involves operationally defined terms which are not available in T', then terms
crucially important in T are not definable within T' and consequently, the laws
of T could not be inferred from those
ofT'.
However, Sperry's holistic approach and the causal stratification of phys-
ical and psychological realities by means of the part-whole relation raises
new questions, which I can only survey here briefly. If mental events are
'emergent' properties of some 'configurational' cerebral events (while the
latter are, according to Sperry, unknown as yet), then, from our logical point
of view, which implies the 'Godel-Tarski theory of logical types', an event is
not literally a property of physical object exemplified by a con-figurational
part of the cortex. A physical or mental event E happens in (or to) the
configurational part of the cortex P in the spatio-temporal region
R if, within R, the part P displays the physical or mental feature F. Hence E
is a triadic relation, involving P, Rand F. Since the ultimate parts of P are
elementary particles and the part-whole relationship obtains between entities of
the same logical rank, we have to ascribe to Prank 4, to R rank 1, to Frank 5.
Hence, the event E, consisting in a triadic relation between P, Rand F has a
much higher logical rank. This does not detract from its reality, nor from the
reality of its causal relatedness to other events.
We have already referred above to Rechtschaffen's work on sleep and
dreams. The philosophical significance of his findings does not consist in a
substantiation of the Cartesian tenet, "anima semper cogitat." No claim is
made that every sleeping episode is accompanied by dreams. However, a
new empirical avenue of approach to a scientific exploration of the sum total
of man's mental experiences is open. Human psychology is not confined to
people who are awake. Psychology is substantially extended by a successful
SUPPLEMENT 265

investigation into dream experiences and the principle of psychophysical


parallelism is strengthened by this extension.
The endocrinological findings, associated with the work of D. Ingle and
others, are adequately summarized in the N. Kleitman volume. They are relevant
to the present status of the principle of psychophysical parallelism for the
follOWing reasons: (1) they extend the empirical basis of parallelism in another
important direction, (2) they seem to justify the more general assumption that
every mental experience has a physiological substratum rather than the more
specific hypothesis providing each experience with a cerebral substratum (3)
they partly overlap with research concerning sleep and dreams since there are
also endocrine theories of sleep.
The relevance of the above three types of empirical advances to the issue of
psychophysical parallelism in the context of the problem of time amounts to the
fact that, on the whole, these advances strengthen the view which claims the
existence of a physiological - although not necessarily neural - substratum for
every mental experience of any human individual. This is at variance with
significant views held, for example, by Bergson. Penfield's findings concerning
the sub-cortical 'world-center' clearly refute this Bergson-ian thesis. The
endocrinological fmdings and those related to the physiological substrata of
dreams appreciably broaden the empirical basis of the assumption that every
mental experience has a physiological basis. Granted, the existence of the
physiological basis of any pre-assigned mental experience does not determine the
nature of the relation obtaining between the experience and its substratum. This
relation may be strict identity as defined by Leibniz (Le.,
A and B are strictly identical if, and only if, they belong to the same classes.
However, the relation may also turn out to be a more subtle synthetic species of
identity, as suggested by H. F eigl. 170 Yet, our only concern, at this juncture, is
laying down a firm foundation for the cosmic time-order which merges
consistently the four above-mentioned types of temporal orders. To this end, the
existence of a bi-unique mapping of the sum-total of the mental experience into a
sub-set of his physiological processes is perfectly sufficient. The existence of this
mapping is certainly rendered more probable by the endocrinological discoveries
in conjunction with findings similar to Penfield's
and Rechtschaffen's.
In the context of this investigation into universal time it is of particular
interest that our intuitive knowledge of time has a reasonably defined cerebral
sustratum. More specifically, our intuitive estimation of duration was shown by
H. von Foerster 171 and H. Hoagland 172 to depend critically upon the oxidative
metabolism of some cells in the brain. Von Foerster used a quantum
266 SUPPLEMENT

mechanical approach to the issue. Hoagland's main theoretical tool is an equation


determining the velocity of chemical reactions first established by Arrhenius on a
phenomenological basis and subsequently derived from chemical kinetics. The
equation involves the following notational stipulations:
(1) v is the velocity of the chemical reaction, (2) T denotes absolute tem-
perature, (3) R is the gas constant, (4) E is the activation energy, i.e., the surplus
energy per mol which the molecules need to react, (5) e is the base of natural
logarithms. By applying these notations, one can formulate the Arrhenius
equation as follows:
E RT
v = constant X e- /
This equation describes the interdependency of the chemical 'pacemaker' (i.e.,
the slowest reaction in a sequence of reactions) and the activation energy to a
high degree of precision. In the case of intuitive time, the value of the activation-
energy can be computed from the effect of the temperature changes on an
estimated timelength. The pacemaker is probably located in the cells of the
reticular formation in the brain stem. The main argument in support of this
cerebral localization amounts to the following: The impulses from the retinal
formation reach the cortex and determine whether sleep or wakefulness prevails
in the relevant individual. Since our ability to keep time persists regardless of
whether we are awake or asleep, its cerebral counterpart should be in a region
which determines both wakefulness and sleep.
The cerebral localizability of our iJ:ttuitive knowledge of time provides a
powerful argument in favor of psychophysical parallelism. The point is that
every mental experience is accompanied by an awareness of its temporal
aspect. If the latter has a cerebral substratum then so has the former. In
conjunction with the fmdings of Penfield, Ingle and Kleitman-Rechtschaffen,
the result of von Foerster and Hoagland seems to provide adequate evidence
in favor of the first component of the principle of psychophysical parallelism.
Finally, I shall try to give a concise account of the methodological
support for the principle of psychophysical parallelism which J. von
Neumann 173 derived from his quantum theory of measurement. This theory
may be summarized as follows: Whenever a measurement of the quantity Ql
is performed on the system SI by using as measuring instrument the system
S2 for which a quantity Q2 is defined, the system S2 is temporarily connected
with SI . As a result of this mensural interaction, changes occur in the rays of
the three Hilbert spaces HI , H2, HI, 2 which represent the quantum states of
SI, S2 and SI, 2 respectively. Complete orthonormal coordinate-systems for these
spaces are denoted, by RJn, R~, R~' 2 respectively. The mensural
SUPPLEMENT 267

interaction of Sl and S2 is successful if, after the interaction is over, the


quantum state of the joint system can be represented by the enumerable
sum of Rl R2 and a complex constant c. If R2 represents an eigenstate of
m n
Q2 associated with the eigenvalue q 1, then the eigenstate of Sl and the
corresponding eigenvalue of Ql are easily computed.
The problem of psychophysical parallelism becomes relevant when the
separation of the measured system Sl and the measuring system S2 is speci-fied.
When the temperature of the wall is measured in terms of the reading of a
mercury thermometer applied to the wall, then the measured object may include
only the wall with its temperature, or, in addition, the mercury column in
conjunction with its momentary column. Moreover, the retina-image of the
thermometer may be added to Sl and even, on top of this, the stimulation in the
optical, occipital region of the brain. According to von Neumann, the
psychological ego of the observer, in spite of being involved in every actual
measurement, must not be included in Sl since, to his mind, a basic
methodological principle of natural science postulates that a physical
measurement must be completely expressible in physical terms.
According to von Neumann this methodological postulate can be met in the
theory of quantum measurement, if, in addition to the principle of
psychophysical parallelism, a theorem of his own is taken into account. This
theorem states that, given a compound system Sl, S2, any 'cut' into the measured
object and the measuring instrument can be effected without violating quantum
mechanical laws. The important theorem of von Neumann's mentioned at the
outset deals precisely with arbitrary shifts of the 'cut'. The question is how much
weight does his reasoning carry with regard to the principle of psychophysical
parallelism?
It must be granted that if the psychological ego of the observer must be
included in the measuring system SZ, then only psychophysical parallelism can
guarantee that the measurement will nevertheless be completely describ-able in
physical terms. From the point of this parallelism, it is obvious that any
psychological state of the ego is uniquely determined by its cerebral substratum.
One may doubt, nevertheless, whether the very definition of a quantum
measurement and the inclusion of a conscious human observer in the
measurement proposed by von Neumann is not unduly anthropomorphic. The
crucial feature of a mensural interaction between Sl with its Ql by means of an S2
with its Q2 is that such a measurement, if performed on a pure ensemble to which
Sl belongs will transform it into a mixture of sub-ensembles, each sub-ensemble
being characterized by a specific value q 1 of Ql. A mensural interaction can
therefore be described without resorting
268 SUPPLEMENT

to psychological or anthropomorphical terms and there is no doubt that most


scientific measurements are performed by automatically recording instruments.
Accordingly, little, if anything, can be gained for the principle of psychophysical
parallelism by resorting to the quantum theory of measure-ment.
SUPPLEMENT

2. CONCEPTUAL ANALYSIS OF PSYCHOPHYSICAL PARALLELISM

1. Introduction

Psychophysical parallelism can be construed as a conjunction of three distinct


psychophysical laws. The first law states, roughly speaking, that the sequence of
consecutive states of consciousness of a human being and the sequence of his
physiological conditions run parallel to each other in time and determine each
other. The second law deals with the causal relations among these two sequences.
More specifically, psychophysical parallelism rejects any causal connection of
the elements of these sequences. The third la'N" claims that the two sequences
are 'essentially' identical: in other words, the entity which sensory perception
views as a physiological event is introspectively a mental event. For the sake of
brevity, I shall refer to the first law as the claim of psychophysical
correspondence, to the second law as the claim of psycho-physical
independence, and to the third law as an assertion of psychophysical identity.

Before embarking on a conceptual analysis of the three claims of psycho-


physical parallelism, I should like to point out what requirements such an
analysis may be fairly assumed to satisfy. A philosophical explanation or
clarification of a claim or a statement, such as the claim of psychophysical
identity, has often been construed, during the last two decades, as amounting to
an attempt to reformulate the statement under consideration in 'ordinary English'.
The philosophical schools of thought, geographically identified with Oxford
philosophers of the mid-forties or mid-fifties of this century would imply such a
view of philosophical clarification or explanation. We have to realize, however,
that such a view is completely out of place when a clarifica-tion of the scientific
uses of language is involved. The scientist uses crucial terms which are not part
of 'ordinary English' and will never become part of it. Thus, it would be
completely wrong to clarify the meaning of crucial logical terms in any multi-
valued logical propositional system in the Oxonian way. Such a clarification
presupposes the availability, within ordinary English, of sentential expressions
synonymous with sentential expressions of a multi-
269
270 SUPPLEMENT

valued logical system. However, the slightest familiarity with these systems
is sufficient for the realization that ordinary English does not contain any
such synonymous sentences.
Accordingly, a clarification of the meaning of term in multi-valued
logical systems cannot and is not obtained by recourse to ordinary language.
The fact is that it can be achieved otherwise. At this juncture, it may suffice
to refer to appropriate textbooks. We have to realize the significance of the
circumstance that the language used by the psychologist, the physiologist, the
neurologist, and the philosopher interested in these scientific disciplines is
not ordinary English. Even the language of Newtonian mechanics, an integral
part of the language of empirical sciences, is not part of ordinary English.
Newton's Second Law of Motion can be expressed in terms of ordinary or
partial differ-ential equations with regard to the temporal variable. But it is
impossible to make sense in ordinary English of a second time-derivative of
spatial coordinates inherent in the Second Law of Motion. Hence, we cannot
expect that a satisfactory, philosophical clarification of the claim of
psychophysical determinations, correspondence, and identity be obtained by
resorting to ordinary linguistic usage. I shall try to obtain such a clarification
by dupli-cating the procedures which have been used successfully in
clarifying me-chanical or logical terms (of multi-valued systems). To some
extent, such a clarification may seem to be circular since, as is well
established in logical semantics and model-theory, the clarification, if
sufficiently comprehensive, must be effected in a 'semantic meta-language' of
the language used by the scientist. It is common knowledge that a semantic
meta-language of an object-language L includes L. On closer analysis,
however, this ostensible circularity proves not to be vicious.

2. Psychophysical Correspondence

The first question which now arises deals with the meaning of the claim that
each mental experience is uniquely 'determined' by its physiological sub-stratum.
We cannot put up with metaphorical formulations often used in the philosophical
literature where we are told, for instance, that the psychological and the
physiological sequences are translations of the same content into two distinct
languages, or that the experience is mirrored by the physiological substratum, is
an equivalent thereof, etc. It is obvious that only by eliminating these ambiguous
metaphors can we come closer to an empirical solution of the mind-body
problem and an understanding of the idea of time underlying
SUPPLEMENT 271

this problem. H. Bergson gives the following explanation of the meaning of


psychophysical correspondence:

A superhuman intelligence, watching the dance of the atoms of which the human brain consists
and possessing the psycho-physiological key, would be able to read, in the work-ing of the
brain, all that is occurring in the corresponding consciousness. 174

This comment contains the assumption that somebody could possibly know
everything about some particular experience. It does not matter that this
assumption is fictitious and unrealizable. What is the meaning of omniscience, or
absolute knowledge with regard to a particular entity? If such omniscience
should consist in the ability 'of reliably making every true statement con-cerning
this entity expressible in some sufficiently comprehensive language, then
obviously omniscience concerning one particular entity would already contain
the omniscience of whatever there is: Leibniz' claim that each monad mirrors the
entire universe is tautologically verified if the claim is translated into the
semantic meta-language. This implies the falsity of the claim of psy-chophysical
correspondence construed ala Bergson. For, should the complete knowledge of
the atomic motions in a brain - or, as this is sometimes put, an 'astronomical
knowledge of the condition of the brain' - yield an absolute knowledge of the
corresponding mental experience and the absolute knowl-edge of a mental
experience provide a knowledge of the entire reality, then the absolute
knowledge of the brain would provide knowledge of the entire reality, if we
confine any empirically available knowledge of the brain: the latter would not
suffice even for the simultaneous events occurring in the immediate
neighborhood of the brain.
Furthermore, it is important to notice that a psychophysical knowledge
construed a la Bergson does not concern the connection between the brain-
condition and the corresponding mental experience but rather the connection
between statements concerning mental experiences and cerebral events.
Accordingly, the connection would be meta-theoretical rather than psycho-
physical. However, the obvious tendency of the followers and the opponents of
psychophysical parallelism is to construe this view as pertaining to the theory,
rather than to the meta-theory of the mind-body problem. Moreover, if the
correspondence relation involved in psychophysical parallelism should uniquely
determine the mental experience connected with a particular substratum, we
would have to admit that this view of the mind-body problem implies a bi-unique
mapping of all the mental experiences of an individual onto an appropriate
subclass of his physiological states. This is certainly not the case in the relevant
formulations of parallelism. They imply a multiplicity
272 SUPPLEMENT

of mental experiences, e.g., auditory sensations agreeing in point of pitch,


timbre, volume, and spatial direction on the assumption that this entire
multiplicity is bi-uniquely mapped onto an appropriate subclass of physiolo-gical
substrata. However, if the psychophysical correspondence relates a variety of
mental experiences to their substrata, we are left in the dark as to the kind of
variety relevant to psychophysical parallelism. In any event, such a partition of
mental experiences and physiological substrata implies that intrinsically similar
experiences can accompany intrinsically similar substrata. Bergson obviously
presupposes that it can be ascertained when mental experi-ences or physiological
processes can be considered as intrinsically similar and maintains that under this
interpretation of adequate similarity (which he refrains from defining),
intrinsically similar physiological substrata need not be accompanied by
intrinsically similar experiences. This interpretation transcends any acceptable
formulation of psychophysical parallelism.
The question arising at this juncture is related to a definition of the set of
those properties which must be shared by intrinsically similar experiences
and substrata respectively. This question can be answered, in the case of
physiological substrata, by stipulating that any two substrata of this type
must not differ from one another when an 'astronomical knowledge' of them
is available, i.e., when they agree with regard to the spatio-temporal distribu-
tion of all basic physical quantities. As pointed out elsewhere in a companion
volume of this writer, such physical quantities can be explicitly listed. To
determine the intrinsic similarity of mental experiences in cases when we are
intuitively inclined to consider them as intrinsically similar, let us consider
the case of two auditory sensations.
We would grant their intrinsic similarity if they did not differ with regard
to volume, pitch, timbre,duration, and, possibly, spatial localization (the last
property has been added in view of stereophonically different auditory
sensations). Thus we have listed only five properties out of the infinite set of
properties attributable, respectively, to these sensations. The five properties
have one feature in common. All of them are gradational.
This suggests that two mental experiences are adequately similar if they share
their gradational properties. Obviously, the gradational property of temporal
location must be excluded since, otherwise, we would have to include
numerically identical experiences in every class of similar experiences provided
they belong in the same stream of consciousness. There may be some doubt
whether this criterion of intrinsic similarity is also applicable to the most
complex experiences. However, this applicability is supported by the fact that the
criterion is obviously applicable to the fundamental simple
SUPPLEMENT 273

experiences, i.e., not only to perceptions and representations but also to


judgments, emotions, and decisions. Thus judgments are gradational with
regard to their most essential feature, viz., the strength of the inherent belief.
There is a continuous transition from firm disbelief, to doubt, assumption, up
to a firm belief. Similarly, emotions are gradational with regard to their
hedonistic feature, i.e., pain or pleasure, or a suitable mixture of both. This
holds also for conscious decisions: there is a continuum of transitions leading
from a firm decision to take a specifiable action to an equally firm decision
to refrain from such action, or to take an alternative course of action. Hence,
if two mental experiences did not differ from one another with regard to all
their gradational properties, there would be no intrinsic difference between
these two experiences with regard to their sensory, representational, emo-
tional and volitional components. It seems pretty obvious that, under these
circumstances, no intrinsic difference could then separate the two mental
experiences.
The following objection might be raised with regard to the proposed
explanation: if I claim that two auditory sensations are intrinsically similar
because they share all their gradational properties, e.g., pitch, I am not merely
saying that each of these sensations has some pitch. I am obviously implying
that they have the same pitch. Consequently, the difficulty concerning the
concept of intrinsic similarity would be merely shifted rather than solved by
resorting to the idea of a gradational property. This objection proves to be
spurious on a closer analysis of the concept of a gradational property.
What is on our mind when we admit that a specific property, e.g., pitch, is
gradational? The following is certainly implied: if two sounds differ with
regard to pitch then there exists a continuous transition between the two
sounds. In other words, for any pair of sounds whose pitch is different there
exists always an intermediary sound which differs less with regard to pitch
from the two initial sounds than the latter differ from each other. If, on the
keyboard of a piano, I strike successively the keys c, d, and e, I shall hear
successively three sounds e, D, and E such that with respect of pitch, D
differs less from e and Ethan e and E differ from one another. This can be
expressed, somewhat figuratively, by stating that, with regard to pitch, Dis
between e and E. The meaning of the betweenness-relation B need not be
spatial although spatial betweenness is a special case of the more general triadic
relation B. If I strike c and e pianissimo and d fortissimo, then B does obtain for
the quartet [e, D, E, pitch] but not for the quartet [e, D, E, volume] . A property
P is gradational if, for at least one trio x, y, z the rela-tion B obtains between x,
y, z and P.
274 SUPPLEMENT

For the sake of brevity ,I shall refer to the particular qualities which pertain to
a single gradational property as the hues o[ this property. The conditions to be
satisfied by the hues of any gradational property can be expressed in terms of the
relation B: (1) if two entities share a single hue of a gradational property P then
these entities do not differ from each other with regard to P;
(2) if an entity £ has a particular hue of a property P and the entity £' does
not differ from £ with regard to P, then £' has the same hue of P. We can
interpret this statement as an implicit definition of the concept of hue. To
make this definition explicit, we state the following: the quality q is a hue of
the gradational property P whenever the ensuing two conditions are satisfied:
(1) two entities sharing the quality q do not differ from one another with
respect to P; (2) if an entity possessed of the quality q does not differ from
another entity with regard to the corresponding property P, then the latter
entity has also the quality q. Thus the pitch p is a hue of the gradational
property Pitch because: (1) any two sounds sharing the pitch p do not differ
with regard to the gradational property Pitch; (2) if some particular sound
has the pitch p and another sound does not differ from the former sound with
regard to Pitch then the latter sound has the pitch p.
It goes without saying that an intrinsic similarity of two physiological
processes and of two mental experiences consists in the fact that the relevant
processes or experiences have the same gradational properties, e.g., they must
agree in point of the pitch of these properties. We therefore conclude that the
claim of psychophysical correspondence should be formulated as follows:
whenever two human organisms 0 and 0' (or rather, the neural and endocrinal
conditions of these organisms) share all their hues, then the accompanying mental
experiences also have the same hues. This gives rise to the objection that it is
extremely improbable that the conditions of two organisms should share all their
hues. In this formulation, the claim of psychophysical parallel-ism would apply
to extremely rare cases, perhaps only to fictitious cases. To rid the idea of
parallelism of this predicament, we have to add a qualification that is tacitly
implied in every lucid formulation of parallelism: it is not only the case that
intrinsically similar experiences accompany intrinsically similar physiological
conditions. In addition, almost similar experiences accompany almost similar
physiological conditions.
This formulation resembles A. Cauchy's defmition of the continuity of
number-valued functions. The French mathematician suggested that a function
[(x) is continuous at the point Xo if [assigns to values of x sufficiently close to
Xo arbitrarily close values of [(x). As a matter of fact, it has been frequent-ly
stated that psychophysical parallelism implies a functional interrelation of
SUPPLEMENT 275

physiological processes and mental experiences with no explanation of this


non-numerical use of the concept of function being given. There is no doubt
that only continuous functions have been taken into consideration. The
alleged fact that gradational properties of experiences are continuous
functions of corresponding gradational properties of their substrata can
consist only in the following condition: there exists a related dyadic relation
obtaining between a gradational property of the substratum and another
gradational property of the experience. The existence of such relations is a
tautological consequence of the idea of relationship and does not express any
empirical feature of the mind-body relation. Psychophysical parallelism
claims more, viz., that sufficiently and intrinsically similar substrata are
accompanied by arbitrarily similar experiences.
How should we explain the relevant concept of continuity? Obviously
intrinsic approximate similarity is not a gradational property. It is not the case
that two objects may possess to a higher or a lower degree the property of
being approximately and intrinsically similar. They may, however, differ
more or less with regard to a particular, gradational property. Accordingly,
we can imagine an infinite sequence of entities (physiological substrata or
mental experiences) whose intrinsic difference, with regard to the relevant
gradational property, keeps decreasing and comes arbitrarily close to nil in all
sufficiently remote elements of the sequence. This will be the case if every
element of the sequence is between any preceding element and the limiting
element (with regard to a specific gradational property) and, moreover, if no
entity is between all the elements of the sequence and the limiting entity. If
such is the case, we shall agree to state that the hue of the limiting entity is
the limit of the aforementioned sequence and the elements of this sequence
converge toward this limit. A function whose domain consists of hues of the
same gradational property is continuous in every convergent sequence of
elements of this domain, and is associated with an infmite sequence of the
values of this function which converge toward the function of the limiting
element. These defmitions show the possibility of formulating, in terms of
the relation B, the following claim of psychophysical correspondence: every
hue of a mental experience of any person is a continuous function of the
sequence of hues of the corresponding sequence of physiological substrata of
this person.
Thus we have obtained an interpretation of the 'determination' of mental
experiences by their physiological substrata, usually referred to in discussions of
psychophysical parallelism. Our formulation is more complex than those
available in the literature of the subject, for instance Bergson's formulation.
276 SUPPLEMENT

However, the proposed formulation comes close to these alternative


formula-tions and is not open to the objections facing the latter formulations.
Our formulation suggests the possibility of different visions of
psychophysical parallelism. Two of these alternative versions are of interest.
They can be referred to as psycho-macro-physical parallelism and psycho-
micro-physical parallelism. These terms are self-explanatory. In view of the
fact that a visual sensation can be caused by the impinging of two light
quanta (in the yellow region of the optical spectrum) on the retina, the
microphysical version seems preferable. The difference between the two
versions can be clarified by the following consideration:
We can subdivide all gradational properties into primitive and derivative
properties. I stipulate that a gradational property P is derivative with regard
to the gradational property pi if the hue of P attributable to any entity is a
continuous function of the hue of pi ascribable to the same entity. One can
show that a set of hues closed under transition (i.e., containing with any hues
H, H' any other hue between them) is conducive to the following definition of a
derivative gradational property: a gradational property is derivative if it is
derivative with regard to at least one gradational property. A primitive property
is, by definition, a gradational property which is not derivative with regard to
any. other gradational property. Derivative properties are exemplified by
statistical properties which depend upon some averaging. For instance,
temperature is a derivative gradational property because it is determined, in a
gaseous mass, by the average kinetic energy of the molecules of this gas. In the
area of mental experiences, the derivative properties can be exemplified by the
melody-pattern of a set of consecutive sounds. This pattern is invariant under
many transformations of the elements of the set.
I have mentioned an argument favorable to microphysical parallelism. But
there are other arguments pointing in the opposite direction. I shall list some of
these arguments because they illustrate the empirical nature of psycho-physical
parallelism. Thus, the subjective appearance of an object visible with the naked
eye seems to be determined by the macrophysical shape of this object. Another
argument may be derived from the law of neural functions often referred to by
the slogan 'everthing or nothing'. The ideas of neural convection which identify it
with a macro physical electrical disturbance provide another argument (Cf.
Lapicque, 'Physiologie generale du systeme nerveux', p. 185, in Nouveau traite
de psychologie by G. Dumas, 1930). An adverse argument can be derived from
Sherrington's claim that, so far, no physiological or histological distinction has
been established between those parts of the brain whose activity is accompanied
by states of consciousness
SUPPLEMENT 277

and other parts of the cerebral system. Hence, since both the microphysical and
the macro physical version of parallelism are susceptible to empirical refutation,
both must be classified under the heading of empirical hypotheses.
The empirical nature of psychophysical parallelism is also strengthened by
the fact that it can be shown to be a consequence of macro physical or micro-
physical determinism. According to strict macro-determinism, every physio-
logical event in a human organism, every moment of the hands, every vibration
of the vocal cords are derivable from any earlier condition of the organism
supplemented by boundary conditions which prevail on the surface of the
organism during the time-interval elapsing between the earlier condition and the
physiological event under consideration. If the organisms of the persons
A and B satisfied the same initial and boundary-conditions during the time-
interval t 1, t 2, then the two organisms, if acted upon by adequately similar
stimuli during the time-interval, would respond in exactly the same way. In
particular, these persons, if asked during this interval about their mental
experiences at the beginning of the interval, would answer at time t2 using the
same words, the same intonation and the same tempo. An observer trying to
reconstruct the mental condition of A and B at time t 1 on the basis of what both
have said at t2 would reach the conclusion that the state of con-sciousness of
both A and B at t 1 was the same. The situation would remain basically the same
if the observer, instead of ascertaining the verbal reactions of A and B at t2 , had
used any other reactions of these persons. He would still conclude that their state
of consciousness was the same at t 1 • If an appropriate experimental arrangement
is resorted to by the observer, he would conclude that their states of
consciousness at t 1 were the same. This is precisely the claim of psychophysical
correspondence.
The situation would change appreciably should human organisms be
governed by statistical rather than by strictly deterministic laws. Similar initial
and boundary-conditions of the organisms of A and B would yield different
states of consciousness at t 2 • However, if the observer examined the behavior of
a 'large' number of persons instead of confining his investigations to just a
couple, he would be able to conclude that the relative frequencies or
incidences of their states of consciousness at t2 are determined by the initial and
the boundary-conditions., The observer would then reach a view of the mind-
body relation which may be called statistical parallelism, in contra-distinction to
deterministic parallelism discussed in the preceding pages.

3. Psychophysical Independence
I shall now discuss briefly the second tenet of psychophysical parallelism
278 SUPPLEMENT

which deals with the causal interrelatedness of mental experiences and physi-
ological processes; Both in daily life and in responsible scientific disciplines
we run into statements about the causal connections of two events, one of
which is a mental experience while the other one is physical. It also happens
that the two causally related events are mental. If I decide to stop thinking
about some specific object and succeed in realizing my decision then few
will hesitate to consider this pair of mental events as causally related. We all
are inclined to admit a case of psychophysical causation in those cases where
the decision to lift a hand is followed by an upward motion of the hand.
The second tenet of psychophysical parallelism depends, for its validity,
upon the defmition of a causal relation which we accept. This issue is dis-
cussed in detail in another chapter of this work where Hume's definition of
causality in terms of regular succession is shown to be untenable, on logical
grounds. Instead of reproducing here the defmition of causality developed in
this chapter, we may assume here that a correct definition of deterministic
. causality must entail the following: if the physical event E is an effect of an
earlier event E', then E' is part of a more comprehensive event E" which is
reproducible with E. The possibility of reproducing E' can be explained as
follows: E' is reproducible with E if, with regard to any event A which does not
differ intrinsically from E, there is an event B intrinsically similar to E' and
partly overlapping with E. The two clauses in this definition are designed to
overcome the objections facing Hume's definition of the causal relation and to
specify which event following upon E is to be considered as an effect
ofE.
The second clause is analogous to the physical tendency which rejects
action at a distance. According to the physical principle, a physical event
E can cause another event E' only by influencing a third event E" which
partially coincides with both E and E'. There is some analogy between this
principle and the fact that mental experiences of a person can directly in-
fluence physical events only if the latter occur in the organism of this person.
Similarly, a direct causal action of one mental experience upon another one
occurs only if both experiences belong in the same stream of consciousness.
This fact suggests the following extension of the concept of reproducibility:
(1) the mental experience A is reproducible with the mental experience B if,
in every person, the occurrence of an intrinsically similar experience A' is
followed by an intrinsically similar experience B'; (2) the mental experience
A is reproducible with the physical event B if every intrinsically similar
experience A' is followed, in the body of the person having the experience A', by
the intrinsically similar physical event B'; (3) more generally, the physical
SUPPLEMENT 279

event A (or the mental experience B) is reproducible with the composite


event consisting of the mental experience C and the physical event D if the
occurrence in the stream of consciousness or in the organism of a specifIable
person of an intrinsically similar composite event [A', B'] is followed in the
stream of consciousness or the organism of everybody by a composite event
intrinsically similar to [A', B'] .
Distinct versions of psychophysical parallelism also correspond to different
types of causal relationship. According to the above definition of the causal
relation, the intrinsic similarity of two causes entails the intrinsic similarity of
their effects. However, in many empirical situations, the emphasis is not put on
the causal dependency of a single hue of some gradational property on the hue of
another gradational property. A causal relation often obtains between two
domains of hues. The subject who responds, in a psychological experi-ment, to
some color, does not react to a single hue but to a domain consisting of
topologically connected hues. In biological cases of causality governed by the
role 'everything or nothing' (e.g., in neural reactions), the effect mater-ializes
whenever the cause belongs in a specifiable domain of hues. The effect fails to
materialize if the alleged cause is outside of this topological domain. Should the
probability-distributions of the mental 'effects' depend upon the single hue, or a
topological domain of the hues of the substratum, then the observer would
conclude that the mind-body relationship is statistical. The parallelism of
physiological conditions and mental experiences may then be termed a statistical
or an indeterministic parallelism.
In every event, under either a deterministic, or an indeterministic inter-
pretation of psychophysical causality, it seems difficult to remain in agree-
ment with the observed facts of either mental or physical nature and deny
their causal interrelatedness. Parallelism can at most be maintained in the
following sense: whenever a mental (or a physical) event has a composite
cause that is partly mental and partly physical, then this mental (or physical)
event also has a single cause which is purely mental (or purely physical). I
shall not discuss any further this remote possibility of substantiating a
version of psychophysical parallelism which consists in the redundancy of
physical (or mental) events in the causal explanation of mental (or of
physical) events. It seems much more reasonable from the standpoint of
scientifIc inductive method, which recommends the reduction to a minimum of
the sum total of assumptions made in a causal explanation of an event, to
surrender the parallelist denial of any genuine psychophysical causal relations.
Plain com-mon sense certainly supports the view that assumptions concerning
the existence of redundant fragments of reality should be eliminated, whenever
280 SUPPLEMENT

possible. If every physical behavior of a human organism were totally ex-


plicable by recourse to purely physical causes then mental causes would be
redundant and no conceivable reason could be adduced in favor of their real
existence.
The ontological and epistemological redundancy of mental experiences would
also result in a similar fashion if a self-contained causal chain of all physical
events were interpreted in an indeterministic rather than a deter-ministic sense.
This would imply, for instance, that the conditional probability of Kant having
written his Critique of Pure Reason in 1757 relative to his purely physiological
initial condition at the start of the writing and his physiological boundary-
condition throughout the entire period of writing is exactly equal to the
conditional probability of his having completed this work relative to a composite
initial and boundary condition with the under-standing that the latter involves
Kant's mental condition during the entire period as an essential ingredient. Let us
point out, however, that on an indeterministic interpretation of causality, it makes
perfectly good sense to assume that the conditional probability of the coming into
being of the Critique of Pure Reason is raised by the additional psychological
assumption relative to his state of mind. The latter would then no longer be
ontologically and epistemologically redundant. This consideration is supported by the
fact that, in spite of all advances in scientific methodology, the only basic method
available for solving scientific problems and for advancing science, the ancient
procedure by trial and error continues to discharge its traditional function.

In general, we have to establish the compatibility of the interrelated


physiological and psychophysical causality. More specifically, we have to make
sure that the physical laws concerning the conservation of energy and
momentum in any temporally extended physiological process are compatible
with the aforementioned psychophysical interference. This may not raise
comparable difficulties in all the relevant cases. Thus Descartes' claim that
mental experiences influence physiological processes (in the pituitary gland)
without upsetting the speed and the kinetic energy of the relevant, organic
molecules was refuted by Leibniz's hint that the conservation of momentum
would be incompatible with a non-redundant psychophysical interference. The
macro physical nature of the case discussed by Descartes and Leibniz is not
relevant here since the conservation of energy and momentum hold within and
without quantum theory. However, if the decision of a human being to lift his
arm were indeterministically connected with the subsequent movement of the
arm, then Leibniz' difficulty would not arise. To show this, it suffices to point out
cases where the arm of a human being is not in an eigenstate of
SUPPLEMENT 281

the four-vector energy-momentum. The decision of lifting the arm would then
modify the conditional probability of an energy-momentum distribution without
violating the two relevant conservation laws.

4. Psychophysical Identity
The claim of psychophysical identity can be construed either literally or in some
figurative metaphorical sense. On the literal interpretation, the claim would
amount to the requirement that the two statements of numerical identity hold: the
mental experience M provided of a physiological substratum
S is numerically identical with S: S is M and M is S. Should a figurative inter-
pretation be suggested, then M would no longer be numerically identical with S
but merely 'essentially' coincide with S. Since I am not familiar with any
satisfactory definition of an 'essential' identity, I shall confme myself, in the
ensuing remarks, to numerical identity of experience and substratum.
Ostensibly the' meaningfulness and empirical decidability of the claim of
literal identity of mental experience and its substratum raise n0 difficulty. For it
would seem, at first sight, that since statements attributing to an experience M or
a substratum S several specifiable properties are frequently decidable by
recourse to observational evidence, this should also hold for the
numerical identity of M and S, i.e., the claim that M and S share all their
properties. In other words, the claim of the numerical identity of M and S would
be empirically meaningful since it has empirically meaningful implica-tions. This
however, even if granted, would at most imply the possibility of empirically
refuting the claim of psychophysical identity rather than em-pirically
substantiating it. The mere observational falsifiability of a statement does not
suffice to ensure its empirical meaningfulness. For when a statement S is
refutable by recourse to appropriate observational results, then so is the
conjunction of S and of any other statement S', even if S' is neither provable nor
refutable on observational or merely logical grounds.
The obvious way to deal with this difficulty is to explore the cases of
successful empirical identifications. If a fmite sequence of independent prop-
erties PN (where N is any natural number) can be shown to be attributable to the
entities E and E:, then we may be in a position to consider the set of N
independent properties as a representative sample of all conceivable properties
and derive therefrom an inductive justification of the numerical identity of E an
E'. However, the inductive inference conducive from a favorable sample of the
population of all the instances of the statement asserting the numerical identity of
two objects holds only if the remaining properties, not included in
282 SUPPLEMENT

the set of instances which constitute the premise of the inductive inference, can
be meaningfully applied to the two objects under consideration.
This obvious condition is not fulfilled with regard to psychophysical iden-
tification. Suppose that the mental experience is the sensation of a simple sound
(without overtones) and that the physiological substratum is the modification of
the auditory cerebral region caused by the neural transmission of the initial
disturbance traceable to the pitch, the timbre, and the volume of the physical
vibration. We have to admit that the cerebral disturbance is not a mere
reproduction of the three attributes of a mechanical vibration. And we have to
consider the relevance of all presently available quantum theories to the
description of the cerebral disturbance in view of the fact that two light quanta
impinging upon a human retina can bring about a sensory experience. However,
as pointed out in another chapter of this work, both a physically meaningful
corpuscular or undulatory description of this disturbance can be formulated in
terms of a finite and, as a matter of fact, restricted set of physical observables
attributable to the cerebral region under consideration or, equivalently, of the
state-function determining the quantum state of this region.

Let us consider the latter possibility since it can be handled more concisely.
The quantum state of a cerebral region, consisting of a fmite number P of
molecules, or of the corresponding finite number Q of the atoms which con-
stitute the molecules or, finally and equivalently again, of the finite number of
elementary particles which, respectively, constitute these atoms, is admit-tedly
unambiguously describable by what Dirac calls a complete compatible set of
observables. The number of these observables is known, unexpectedly small and
includes such items as the eigenvalues of the positional operator, of the energy-
momentum operator, of the regular spin, the isotopic spin, the generalized charge
(electric, leptonic, or baryonic) etc. If we select anyone of these observables, we
shall easily find out that no eigenvalue of the selected observable is meaningfully
attributable or deniable to the mental experience involved. The only conclusion
which follows from this fact is that the concept of numerical identity becomes
essentially vague when applied to a psycho-physical pair of events.

In this event, a result of an earlier investigation of mine is utilized. This result


asserts the vagueness of every empirical term but claims that the vague-ness
deprives the sentential context of the term of a definite truth-value only under
specifiable conditions. Thus, both the term USA and North America are vague
because both are, respectively, applicable to various pairs of geo-graphical
entities, without violating the meaning attached to these terms in
SUPPLEMENT 283

some specific language. For instance, both the size of the USA and of North
America are vague because distinct values sufficiently close to each other
can be ascribed to the country and the sub-continent without violating the
correct interpretation of the two geographical terms. These two terms will be
said to be essentially vague in a given sentential context if the truth-value of
this context is affected in various ways by the selection of two admissible
inter-pretations of the terms. For instance, the statement that sizes of the
geogra-phical areas are both expressible in terms of rational numbers,
relative to an arbitrarily chosen area unit, has obviously no definite truth-
value and the two terms are then said to be essentially vague in the context of
this sentence. On the other hand, if a statement involving the two terms has
an invariant truth-value under every admissible interpretation of both terms,
the vagueness of the latter is said to be accidental in this context. This is the
case in the assertion that the USA is part of North America.
What I am driving at is that the vagueness of the concept of numerical
identity is essential rather than accidental when applied to a psychophysical
pair of events. An examination of a few typical properties of elementary
particles will easily show the reason for this claim. However, the concepts of
linear or angular momentum, of rest-mass, of strangeness are simply
undefined with regard to the class of all mental experiences. The result of
this situation is that either attributing or denying anyone of these concepts to
a mental experience is a meaningless sequence of terms, just as the sequence
'either, not, two' does not constitute a meaningful sentence and is,
accordingly, devoid of any specific truth-value.
We have to conclude, accordingly, that the concept of numerical identity
is essentially vague in psychophysical contexts and that their tenet of psy-
chophysical parallelism has no truth-value either, on a closer analysis. Our
misgivings concerning the second tenet of parallelism have been formulated
in subsection 2. Our over-all conclusion reads therefore as follows: the
second parallelist claim, denying that any causal relations obtain between the
two members of a psychophysical pair, does not stand up under a closer
examina-tion, regardless of whether causality is construed in a deterministic
or an indeterministic way. The parallelist assertion is merely a pseudo-
assertion. We are therefore left only with the first parallelist tenet which
claims the existence of a physiological substratum in the organism of a
human being having a specific mental experience.
The main objective of the above analysis, viz., securing a firm foundation
for one cosmic time which consistently merges the four types of temporal
order, is already secured by the first tenet and strengthened by significant
284 SUPPLEMENT

fmdings of Ingle, Penfield, Rechtschaffen, and their associates during the last
decades. By isolating the empirical basis for the existence of a single cosmic
type in this way, we duplicate, in a sense, our exploration of the reality of
physical time in relativistic theories involving second quantization, e.g.,
quantum electrodynamics. An analysis of this theory shows that the time-
variable is a c-number and not a q-number in this up-to-date theory. All the
objections to the physical reality of measurable quantities represented in the
formalism of the relevant quantum theory, by a q-number (Le., a Hermitean
operator satisfying some well-known conditions) are therefore inapplicable
to time in quantum electrodynamical contexts. In both cases, I have applied
basically the same procedure: in order to solve a problem of ontology or
epistemology related to a physical, or a composite, psychophysical concept, I
had to explore the meta-theory of the relevant theory. If this meta-theory is
set up in a sufficient comprehensive fashion, then most philosophical
(ontological and epistemological) issues raised by the theory turn out to be
solvable in an appropriate meta-theory.

5. Relevance of an Empirically Reinforced Parallelism to the Status of


Cosmic Time

In the opening sections of this chapter, I have emphasized the dependency of


cosmic time on the validity of the principle of psychophysical parallelism and
have commented that, in order to provide a firm foundation for cosmic time the
least speculative version of parallelism should prove sufficient. This non-
speculative version claimed the existence of physiological substrata only for
sensory and motor experiences. However, the validity of the laws of cosmic time
was implicitly surrendered for the decisive class of mental experiences which
involve neither sensory nor motor components. This restriction has now become
unnecessary, owing to findings similar to Ingle's, Penfield's, and Rechtschaffen's.
The mental experiences occurring in dreams of the REM and other types are
accordingly provided with a specifiable location in cosmic time. This holds also
for those mental experiences which do not involve any sensory or motor
components. In the course of problem solving behavior, certainly the most
characteristic feature of man, there need not be any sensory or motor experience
at all. A distinguished British historian of philosophy once shared a departmental
office with me. He asked that our office hours not overlap because he talks aloud
whenever he does some thinking. This feature is certainly exceptional. So is the
occurrence of talking in one's head
SUPPLEMENT 285

during the process of problem solving or of performing this activity in some


specific language.
This last type of extra-linguistic problem solving behavior may be illus-
trated by an incident in John von Neumann's intellectual life. He accepted an
invitation from Z. Steinhaus to take part in an automobile drive, took the
steering wheel and was asked by his Polish host to solve a pretty tough
mathematical problem while driving the car. The standard solution would
have required the summation of a complicated hypergeometric series. There
was a short-cut enabling one to dispense with tackling a hypergeometric
series. While carefully handling the steering wheel, von Neumann was asked
by Steinhaus, after the car had come to a complete stop, whether or not he
had made use of the short-cut to get the correct solution. The Princeton
mathematician was obviously surprised by the sort of question that had been
asked of him and answered sincerely that he had no memory of how he
solved it. If he had talked in his head using some specific language, the
memory of the remembered words and sentences would, in all probability,
have enabled him to say how he had done it. Von Neumann's speech-free
mathe-matical argument acquires a specific location in cosmic time owing to
Pen-field's fmdings.
NOTES

* [For complete bibliographical information regarding the pertinent works of the


philosophers and scientists mentioned in this Introduction, see the general bibliography at
the end of Volume H.-Ed.]
1 The problem of terminology is particularly delicate. We have decided to use modern
terminology almost throughout, except to indicate classical equivalents where necessary. This
has allowed us to emphasize better the continuity between the classical form of the
causal theory and its more recent form.
2 Gesammelte Werke. 3rd series: Leibnizens mathematische Schriften, ed. by C. I.
Gerhardt. 7 vols. Halle, H. W. Schmidt, 1849-63. See 'Initia rerum mathematicarum
metaphysica', vol. 7, pp. 17-29. ['The metaphysical foundations of mathematics'] (partial
translation) in Leibniz' Philosophical Papers and Letters, translated and edited by Leroy E.
Loemker. 2nd ed. Dordrecht, Reidel, 1969, pp. 666-74.] Referred to
hereafter as 'Metaphysical foundations'.
3 J. Baumann, Die Lehren von Raum, Zeit und Mathematik in der neueren Philosophie.
2 vols. Berlin, G. Raumer, 1868-69, p. 93.
4 E. Cassirer, Leibniz' System in seinen wissenschaftlichen Grundlagen, Marburg, N.
G. Elwert, 1902, p. 282. (Referred to hereafter as Leibniz.)
5 H. Reichenbach, 'Die Bewegungslehre bei Newton, Leibniz u. Huyghens', Kantstudien
29 (1924),416-438. Translated as 'The Theory of Motion according to Newton, Leibniz
and Huyghens' in H. Reichenbach Selected Writings, 1909-1953. Ed. by Maria Reich-
enbach and Robert S. Cohen. 2 vols. Vienna arcle Collection, vol. 4. Dordrecht, Boston,
Reidel, 1978. Vol. 2, pp. 48-68. Cf. H. Bergmann, Der Kampfum das Kausalgesetz in
der jiingsten Physik. Braunschweig, F. Vieweg, 1929, p. 11. Translated in part as 'The
Controversy concerning the Law of Causality in Contemporary Physics' in Logical and
Epistemological Studies in Contemporary Physics. Boston Studies in the
Philosophy of Science, Vol. 13. Dordrecht, Reidel, 1974, pp. 395-462. Referred to hereafter
as 'Controversy'.
6 Cf. H. Eibl, 'Das Problem der Zeit bei den antiken Denkern', Archiv [iir systematische
Philosophie, (1922-23).
7Cassirer, Leibniz, p. 90.
S Cf. E. Aster, Geschichte der neueren Erkenntnistheorie, Berlin, Leipzig, de Gruyter,
1921, p. 266.
9 Cf. E. van Biema, L 'espace et Ie temps chez Leibniz et chez Kant, Paris, Alcan,
1908. (Referred to hereafter as L 'espace.)
10 Cf. Leibniz' correspondence with Clarke, third letter, translated in Leibniz' Phi-
losophical Papers and Letters. Translated and edited by Leroy E. Loemker. 2nd ed.
Dordrecht, Reidel, 1969, p. 683.
11 Notice however that this practically insignificant difference corresponds to a notable
epistemological superiority of the r(flational theory. The substantialist theory assumes the
existence of two fundamental classes of entities, instants and events, the latter owing their
existence to the former. In the relational theory there is only one fundamental

286
NOTES 287

class, that of events, from whose properties the definition and existence of the instants can be
derived. Leibniz's theory satisfies only this postulate: "entia praeter necessitatem non sunt
multiplicanda" [entities must not be increased beyond necessity] - and in this it makes decisive
progress over the substantialist theory.
12 ..• "Ie present est toujours gros de l'avenir, et ... aucun (!tat donne n'est explicable
naturellement, qu'au moyen de celui, dont il a ete precede immediatement." From a letter,
probably addressed to Varignon. G. W. Leibniz, Hauptschriften zur Grundlegung der
Philosophie, ed. by E. Cassirer. 3rd enl. ed. 2 vols. Hamburg, Meiner, 1966. vol. 2, p. 557.

13 Leibniz, 'Metaphysical foundations', p. 666.


14 Ibid.
15 Ibid.
16 Cf. L. Couturat, La logique de Leibniz, Paris, 1901, p. 73.
17 Van Biema, L 'espace.
18 This appears even in Kant although he supported the idea of absolute time:
"Now since absolute time is not an object of perception, this determination of posi-
tion cannot be derived from the relation of appearances to it. On the contrary, the
appearances must det~rmine for one another their position in time, and make their
time-order a necessary order. "
Critique of Pure Reason. Translated by Norman Kemp Smith. New York, St. Martin's, 1961,
p. 226. (Referred to hereafter as Critique.)
19 Ibid., pp. 224-25.
20 Compare with this passage the following remark of Leibniz:
"But we know as coexisting, not merely those things which we perceive together, but also
those which we perceive successively, provided only that, during the transition from the
perception of one to that of the other, the former is not destroyed and the latter generated."
['Metaphysical foundations', p. 671.]
21 Kant, Critique, pp. 233-34.
22 Ibid., p. 228.
23 A. Schopenhauer, On the Four-fold Root of the Principle of Sufficient Reason.
Translated by E. F. J. Payne. La Salle, Ill., Open Court, 1974, p. 131. (Referred to hereafter as
Four-fold Root.)
24 Kant, Critique, p. 237.
25 Cf. A Drews, Kants Naturphilosophie als Grundlage seines Systems, Berlin,
Mitscher and Rostell, 1894.
26 This depends on his definition of simultaneity in terms of interaction.
27 Schopenhauer, Four-fold Root, p. 131.
28 Kant, Critique, pp. 285 -286.
29 Schopenhauer, Four-fold Root, p. 131.
30 Kant, Critique, p. 237.
31 "La dependance representee comme rapport de duree ne nous offre autre chose que Ie
rapport meme ex prime par Ie principe de causalite."

(The dependence represented as the relation of duration gives us nothing but the relation itself
expressed in terms of the principle of causality.)
Philosophie fondamentale, French translation by Manec. (Paris, 1852. Book X, Chapter VI,
no. 82) from Filosofia fUndamental, 2nd ed. Barcelona, Brusi, 1848.
288 NOTES

32 "It is only in the actual content of what happens, not in a form present outside it into which
it may fall, that the reason can be found for its elements being related to each other in an order
of succession ... What we call Past, we regard primarily as the condi-tion 'sine qua non' of the
Present, and in the Present we see the necessary condition of the Future. This one-sided relation
of dependence, abstracted from the content so related and extended over all cases which it in its
nature admits of, leads to the idea of an infinite Time, in which every point of the Past forms the
point of transition to Present and Future, but no point of Present and Future forms a point of
transition to the Past."

Metaphysic. Ed. by Bernard Bosanquet. 2nd ed. 2 vols. Oxford, Clarendon Press, 1887.
Vol. I, p. 334 and 336.
33 Paris, Alcan, 1896. (Referred to hereafter as Etude.)
34 Lechalas was the first to speak of a 'causal theory of time'.
35 Lechalas, Etude, p. 170.
36 Ibid., p. 176.
37 Lechalas ingeniously shows that continuous time can be constructed while completely
maintaining the hypothesis, necessary according to him, that all changes are discontinu-
ous, being formed of a finite number of consecutive states (Lechalas was an advocate of
the so-called "law of number" of Renouvier and Pillon). This is reached by assuming that
for every given change there is a corresponding change, such that two states of this second
change, while simultaneous with two consecutive states of the first change, respectively,
are not consecutive themselves. This construction, invoked by Lechalas to support the
causal theory, is possible in every relational theory of time as well.
38 It should be noted that the notion of configuration assumes a method which allows one to
recognize two states as belonging to the same point. In Chapter VI, we will see, in discussing
Russell's theory, how the principle of this method can be formulated in
terms of the causal relation; this method amounts to choosing all the sets of the States
E'J: which can be obtained by varying only the number m.
39 The relations between the different causal definitions of the temporal order can be illustrated
by repeating the diagram used in connection with Leibniz (p. 47):

leibniz Kant Lechalas Lechalas-Kant Causal


Decompositions
effect effect

0
~ 0
~ 0
~

cause cause

1 u

simultaneity simultaneity simultaneity simultaneity simultaneity


interaction geometry gravitation causal
canfi gura ti on

Fig. 3.
NOTES 289

40 "Die klassische Feldphysik behauptet, dass man die physikalische Wirklichkeit bes-
chreiben kann ... indem in einem vierdimensionalen Raum-Zeitgebiet fUr jeden Punkt
gewisse messbare Grossen - Feldstiirke, Gravitationspotentiale u. s. w. - zahlenmiissig
angegeben werden. Und dabei besteht eine Kausalitiit im folgenden Sinne: denken wir uns
ein endliches Stuck des Raumes abgegrenzt, etwa in der Form eines Kastens ... Zu einer
bestimmten Zeit, - sagen wir urn 11 Uhr - moge der physikalische Zustand inner-
halb des ganzen Kastens vollstandig ausgemessen sein. Ferner soIl der physikalische
Zustand auf der ganzen Oberfliiche des Kastens von 11 Uhr bis 12 Uhr dauernd kontrol-
liert werden. Durch die so festgestellten Umstiinde sind die physikalischen
Vorgiinge im Innern des ganzen Kastens von 11 bis 12 Uhr eindeutig bestimmt.
Reproduziert man zu beliebiger Zeit am beliebigen Orte den Anfangszustand des Kastens und
den zeitlichen Verlauf der Vorgiinge an seiner Oberfliiche, so reproduzieren sich von selbst alle
Vor-giinge im Innern des Kastens." [Die Naturwissenschaften, A/II, 1927, p. lOS.]
41 " ••. Die Quantenmechanik beschreibt die Welt mit Hilfe eines abstrakten·Koordina-
tenraumes, der ungeheuer viele Dimensionen besitzt: die Anzahl der Dirnensionen ist
proportional der Anzahl aIler in der Welt vorhandenen Materieteilchen. In diesem ab-strakten
Raum bewegen sich wiederum stetig ausgebreitete Grossen, die aber nicht unmittelbar das
Einzelgeschehen in der atomaren Erscheinungswelt beschreiben, sondern nur die
Wahrscheinlichkeit quantenhafter Prozesse bestirnmen. Die Kausalitiit - nicht als ein
metaphysischer Gegensatz zu einem metaphysis chen Zufallsbegriff aufgefasst, sondern als die
friiher formulierte physikalische Aussage verstanden - gilt in formal
vollig gleicher Weise fiir beide Theorien." [Ibid.)
42 The formulation assumes that all of becoming is given en bloc, contrary to indeter-minist
tendencies which nowadays very clearly assert that the past can be considered only as 'unfolded'
(Bergson), 'existing' (Kotarbinski, Broad), 'determined' (Reichenbach), as opposed to the future
which is 'unforeseeable', 'nonexistent', 'undetermined'. In order to avoid countering these
tendencies, the entire future, starting with an arbitrarily chosen instant can be excluded from the
perception of this being without changing anything essential in the above hypothesis.

43 It will be noted that in the general case the notion of the instantaneous state of a
material point is replaced with that of the event, that is, with the instantaneous state of a
spatial point (which cannot coincide with a material point). The notion of 'confIgUration'
then becomes superfluous.
44 Lechalas, Etude, p. 174.
45 Russell, The Analysis of Matter. London, K. Paul, Trench, Trubner, 1927, p. 38l. (Referred
to hereafter as Analysis.)
46 In the example of the chicken and the egg cited by F. Enriques (cf. Les problemes de
la science et de la logique, translated from the Italian [Problemi della scienza. Bologna,
Zanichelli, 1906.] by J. Dubois, Paris, 1909, p. 207; English translation, Problems of
Science, by K. Royce, Chicago, Open Court, 1914), it certainly seems that the effect can, in its
turn, become the cause of its cause. But the biological process under consider-ation here is
obviously not reversible in the sense explained above, that is, it does not decompose into partial
processes capable of being run equally in both directions. This example also shows that if all
causal relations were symmetrical in a certain set of phenomena, the processes there would be
cyclical. Mechanical laws are all reversible
although mechanical processes are not always cyclical: hence, the symmetry of the - causal
relation is not equivalent to the reversibility of phenomena.
290 NOTES

47 Note, however, that there is a problem here that cannot be discussed within the framework
of the present study. We have just stated that two physically interchangeable systems are not so
in every respect. But what prevents us from believing that the two systems are really only one,
present at two points in space? Undoubtedly this ubiquity is surprising and runs counter to our
instinctive principle of individualization, but it is not at all contradictory (as careful analysis
would prove). In other respects it is much less surprising today than it was some years ago,
before wave mechanics and the new statistics of Bose and Fermi. The difficult problem of the
physical individual, in which lies hidden the last irrational trace of sensible experience, would
require detailed examination.
48 Cf. the interesting article of Edgar Zilsel who limited himself to deducing the aniso-tropy of
time from the law of the increase of entropy in 'lJber die Asymmetrie der
Kausalitiit und die Einsinnigkeit der Zeit', Die Naturwissenschaften, 1927.
49 Lechalas,Etude, p. 176.
50 Cf. E. Husserl, Vorlesungen zur Phiinomenologie des inneren Zeitbewusstseins, Halle,
Niemeyer, 1928. (The Phenomenology of Internal Time-Consciousness. Edited by
Martin Heidegger, translated by J. S. Churchill. Bloomington, Indiana University Press, 1964.
Referred to hereafter as Internal Time·Consciousness.)
51 H. Bergson, Essai sur les donnees immediates de la conscience, Paris, Alcan, 1889.
(Time and Free Will, an Essay on the Immediate Data of Consciousness, trans. by
F. L. Pogson. London, Allen and Unwin, 1971.)
52 [The 'old Medea', Jason's wife (and Aeson's daughter-in-law), rejuvenated the almost dying
Aeson at Jason's request by giving him a very efficacious magic potion to drink. cf.
Ovid, Metamorphoses, VII, 162.]
53 Lechalas, Etude, pp. 184-185.
54 A. A. Robb, A Theory of Time and Space, Cambridge, Cambridge University Press,
1914.)
55 I shall enumerate only the first four axioms.
56 Carnap, 'Ober die Abhiingigkeit der Eigenschaften des Raumes von denen der Zeit',
Kantstudien, 1925. (To appear in an English translation as 'On the dependence of the
properties of space upon those of time' in Carnap's Essays in Philosophy of Science,
1921-1928. Ed. by Arthur Benson, 2 vols. Vienna Circle Collection. Dordrecht, Boston,
Reidel. Forthcoming 1981.)
57 Abriss der Logistik, Vienna, Springer, 1929. The sections of this precis which con-cern the
causal theory are only a summary of an unpublished work which Carnap gladly
placed at my disposal. I wish to thank him for it here as well as for the further explana-tions
which he gave me. Referred to hereafter as Abriss.
58 However, cf., Bergmann, Kampf, p. 51.
59 Two contemporary philosophers have questioned the legitimacy of this passage to the limit,
underlining the fact that the scientist also works with data of finite dimensions and that all the
statements of science with apparent bearing on world points, can and must be interpreted as
convenient and abbreviated ways of speaking whose real meaning, in the final analyses,
concerns events extended in space and time. In setting this epistem-ological task for themselves,
Russell and Whitehead had wanted to bridge the gap which seems to separate the universe of
science, made up of world points, from the immediate data of experience and to re-establish
continuity between them. I will not examine the consequences concerning the choice of
indefinable terms in the axiomatization of the causal theory of time which can be derived from
the result of Russell's and Whitehead's
NOTES 291

researches. This result is only relative, for while the authors did in a certain sense shorten
the gap between the perceived and unperceived, they did not succeed in bridging it
completely. Note only that the singular complexity of the formulae which the authors
introduced contrasts very sharply with the comparative simplicity of the systems of Robb
and Carnap.
60 Cf. for example, A. Tarski, Fundamentale Begriffe der Methodologie der
deduktiven Wissenschaften, Leipzig, 1930, p. 9.
61 Let us recall only the 'extension' and 'cogredience' in terms of which Whitehead tried
to define the conceptual apparatus of physical theory, the 'co-punctuality' of Russell,
the 'genidentity' of Lewin, the coincidence and real point of Reichenbach.
62 H. Reichenbach, Axiomatik der relativistischen Raum-Zeitlehre [Die Wissenschaft, 72],
Braunschweig, Vieweg, 1924 (Axiomatization of the Theory of Relativity, tr. by M.
Reichenbach. Berkeley, University of California Press, 1969); 'Die Kausalstruktur der Welt und
der Unterschied zwischen Vergangenheit und Zukunft', Sitzungsberichte, Bayerische
Akademie der Wissenschaften (Nov. 1925), pp. 133-175 (,The Causal Struc-ture of the World
and the Difference between Past and Future' in H. Reichenbach Se-lected Writings, 1909-
1953. Ed. by Maria Reichenbach and Robert S. Cohen. Vienna Grcle Collection, volume 4.
2 volumes. Dordrecht, Boston, Reidel, 1978. Volume 2, pp. 81-119.

63 K. Lewin proposed an essentially identical method in his article, 'Die zeitliche


Geneseordnung', Zeitschrift fiir Physik 63, 13.
64 Cf. A. Einstein: 'Ueber die spezielle und die allgemeine Relativitatstheorie. Braunsch-
weig, Frieder, 1917, p. 15. (Relativity: the Special and General Theory. trans. by Robert W.
Lawson. London, Methuen, 1920.) In his 1905 paper, Einstein gave a very similar
definition, where spatial metrics is not used. I will return to this on page 124.
6S The choice of a light ray as messenger is of particular relevance with regard to the role it
plays in Relativity.
66 A similar idea can be found in an article by Bass, entitled 'Kausalgesetz und Zeitrich-
tung', Annalen der Philosophie 6.
67 A more natural example is perhaps the movement of the shadow of an illuminated object on a
sufficiently long screen; the sequence of day and night is also an umeal one. 68 Cf., for example,
A. Meinong, 'Zum Erweis des allgemeinen Kausalgesetzes,' Sitzungs-berichte der Wiener
Akademie, Philosophisch-historische Klasse 189, 4. Abhandlung (1918).

69 H. Bergson, Duree et simultaneite. Paris, Alcan, 1926, p. 116. (Duration and Simul-
taneity, translated by Leon Jacobson. Indianapolis, Bobbs-Merrill, 1965.)
70 R. Carnap,Abriss, Sections 36-37. .
71 Notice, however, that Hume's definition of the causal relation, just like those of Carnap
and Reichenbach, is also open to criticism and seems indefensible in the present state of
philosophy, as I shall try to prove in the second part of this essay. Similarly, Kant's opinion,
according to which causality is a category irreducible in terms of the understanding, seems
scarcely defensible to anyone who identifies every irreducible
category with the indefmable notions of pure logic. It seems to me that a thorough analysis
of causality is absolutely indispensable for the causal theory of time and it will be seen in
the second part of this essay that it leads to a definition of the causal relation which does
not presuppose the notion of time.
72 B. Russell,Analysis, p. 102.
292 NOTES

73 Cf. P. W. Bridgman, The Logic ofModem Physics, New York, Macmillan, 1927, p. 5.
et passim.
74 Here the point marks the simultaneous realization, the 'logical product'.
75 Lechalas, Etude, p. 170.
76 Cf. Bergmann, Kampf, p. 21.
77 Cf. M. Smoluchowski, 'Giiltigkeitsgrenzendes zweiten Hauptsatzes der
Wiirmetheorie'. Oeuvres de Marie Smoluchowski, Cracow, 1927, Vol. 2, p. 361-398.
(Referred to here-after as 'Giiltigkeitsgrenzen'.)
78 Cf. Eddington, The Nature of the Physicol World. Cambridge, Cambridge University
Press, 192 8.
79 Smoluchowski, 'Giiltigkeitsgrenzen', p. 392.
80 We saw in connection with Lechalas that the notion of simultaneity can be defined in
terms of the method of causal decompositions even in the hypothesis of universal
reversibility.
81 Here is an exact, though slightly forbidding formulation of this definition: "event B
takes place 'between' A and C, if the class of effects of A (C), contemporaneous with C(A)
is identical to the class of effects contemporaneous with C(A) produced by the effects of
A(C) contemporaneous with B, while the class of effects of B, contemporane-ous with
C(A) is not identical to the class of effects contemporaneous with C(A) and produced by
effects of B contemporaneous with A (C).' In this definition the effect of event X is any
event related to X by symmetrical causality, that is, any event which is either a cause or
effect (in the ordinary sense of the word) of X.
82 "Denken wir ... es sei ein Film gegeben, auf welchem ein Stein zur Erde fliegt, ein Bach
vom Berge herunterfliesst usw. Wir wussten nicht, wie wir den Film aufrollen sollen. Wird
der Film in einer Richtung aufgerollt, so scheint sich der Stein zu erheben, der Bach fliesst
den Berg hinauf; wird er in der anderen Weise aufgerollt, so sehen wir das Gegenteil. Urn
nun die Aufrollung des Films, d. h. die Richtung Vergangenheit-Zukunft festzulegen,
miissen wir ein Kausalgesetz zugrunde legen, und zwar brauchen wir nicht ein besonderes
fUr den Stein und ein besonderes fUr das Wasser, vielmehr geniigt es, ein Kausalgesetz
willkiirlich festzusetzen, z. B.: Das Wasser fliesst bergab, dann ist uns die Richtung aller
anderen Kausalverhiiltnisse auf dem Film erfahrungsgemass gegeben ...
Aber ein Kausalvorgang muss willkiirlich festgesetzt werden ... Wir setzen einen nicht
umkehrbaren Vorgang fest und defmieren was Ursache und was Wirkung ist. Wir de-
finieren flir die Welt: Die Entropie in der Welt nimmt zu. Damit ist die Richtung Vergan-
genheit-Zukunft defmitorisch festgelegt ...." [Bergmann,Kampf, p. 15.]
83 For the explanation of the logistical notions which I was able to avoid in the above
analysis, I refer the reader to the introduction of the Principia Mathematica.
84 Russell,Analysis, p. 123.
85 Ibid., pp. 245-246.
86 [Complete bibliographical information regarding the works of the philosophers
mentioned in this section will be found in the general bibliography at the end of Volume
n.-Ed.]
87 References to Part I of this work will be indicated by 'I', followed by the page number.

88 I shall try to show below that the imperceptible nature of bare time is both shared by
corpuscular objects and compatible with the assignment of the logical rank of individuals
to spatio-temporal regions.
NOTES 293

89 Cf. V. Benussi's monograph, Die Psychologie der Zeitauffassung. Heidelberg,


Winter, 1913; and P. Fraisse, Psychologie du temps. Paris, Presses Universitaires de
France, 1957 (The Psychology of Time, tr. by Jennifer Leith. New York, Harper and Row,
1963; rev. ed. London, Eyre and Spottiswoode, 1964.)
90 The study done by P. Janet on 'temporal conduits' ['conduites temporelles'] can be
included in the same group of problems: L 'evolution de la memoire et la notion du
temps. Paris, Chachine, 1928.
91 Cf. I, p. 50.
92 The argument that the existence of the defining relation presupposes (the existence of) its
terms calls for a further remark. We can undoubtedly define existing relations between non-
existent terms, but these relations possess very special properties, which are
irrelevant to the relation 1inkmg the remembrance to the state it concerns. Thus, the
relation between X and Y, defined by the proposition 'X is not Y', does exist, provided
that X exists, even if Y does not. But this relation has the property of linking X not only to
Y, but also to every Z, provided that the term 'z' has the same extension as 'Y' and
designates, in this case, a non-existent object. If the regular polyhedron X is not a penta-
hedron, neither will it be a heptahedron. But the relation of the memory to the state it
concerns does not have this characteristic property of relations capable of linking non-
existent objects: whoever thinks he remembers a regular pentahedron does not remem-ber
a heptahedron. Therefore, the relation used in definition (I) does presuppose the existence
of its terms.
93 Cf. Hussed, Internal Time-Consciousness, Section 3.
94 This claim is made on the assumption that only dreamless sleep episodes occur in human life.
For questions concerning time experienced in dreams, cf. N. Kleitman, Sleep and
Wakefulness, Chicago, University of Chicago Press, 1939; rev. and enl. ed. 1963;
Rechtschaffen, Wakefulness and Dreams, 1966.
95 Cf. J. Cohen, 'Subjective Time' in Voices of Time, ed. by J. T. Fraser. N. Y., Braziller, 1966,
pp. 257-75; and J. M. Stroud, 'The Fine Structure of Psychological Time' in
Interdisciplinary Perspectives of Time. Annals of the New York Academy of Sciences
138,2 (1967), 623-631.
96 A. N. Whitehead, An Enquiry Concerning the Principles of Human Knowledge. 2nd
ed. Cambridge, Cambridge University Press, 1925. (Referred to hereafter as Enquiry.)
97 Cf. B. Russell, Our Knowledge of the External World. London, Allen and Unwin,
1914. (See 4th Lecture, 3.)
98 H. E. Lehmann, 'Time and Psychopathology' in Interdisciplinary Perspectives of
Time. Annals of the New York Academy of Sciences 138, 2 (1967), 798-821.
99 G. Schaltenbrand, 'Consciousness and Time', ibid., pp. 632-645.
100 J. M. Stroud, 'The Fine Structure of Psychological Time', ibid., pp. 623-631.
101Goldstone, 'The Human Oock', ibid., pp. 767-783.
102H. Hoagland, 'Some Biochemical Considerations of Time' in The Voices of Time.
ed. by J. T. Fraser. N. Y., Braziller, 1966, pp. 312-29.
103Ibid.
104 'Consciousness and Symbolic Processes' in the Fourth Conference on Problems of
Consciousness, Josiah Macy Foundation, 1953.
105Cf. I, p. 91.
106 I believe that Husser! pointed this out in one of his unpublished lectures on the
'problem of constitution'.
294 NOTES

107C. Caratheodory, 'Zur Axiomatik der speziellen Reiativitatstheorie',Sitzung;vberichte


der Preussischen Akademie der Wissenschaften, Phys. Math. Klasse 5 (1924), 15.
108 Cf. I, pp. 93-94.
109 Cf. R. Carnap, 'Ober die Einheitssprache der Wissenschaft' in Actes du Congres
international de philosophie scientifique. Paris, Hermann, 1936. Fasc. 2: Unite de la
science, p. 64.
110 H. Wcyl, Deutsche Literaturzeitung, 1924 (coL
2122). III Math. Annalen 61,1909.
112 P. A. M. Dirac, Principles of Quantum Mechanics, Oxford, Clarendon Press, 1930.
[4th rev. ed., 1974].
113J. v. Neumann, Die mathematischen Grundlagen der Quantenmechanik, Berlin,
Springer, 1932. (Mathematical Foundations of Quantum Mechanics, tr. by R. T.
Beyer. Princeton, Princeton University Press, 1955. Referred to hereafter as Quantum
Me-chanics.)
114 Cf. section 4c. of this chapter.
115 Cf. I, p. 58.
116 Cf. I, pp. 120-123.
117 Cf. I, pp. 110-11; and P. Hertz's article 'Kritische Bemerkungen zu Reichenbachs
Behandlung des Humeschen Problems', in Erkenntnis 6, (1937), pp. 25-31.
118 Generally, there is a causal relation between two events linked by a light or other
signal, even if it seems theoretically or technically impossible to decide which of the two
coincides with the departure or the arrival of the signal.
119 Moreover, it is easy to state, within the system, a sufficient and necessary condition of
the causal relation - invariant non-simultaneity. Cf. p. 214, Theory 1.
120 We use this term here in its current sense, in which every physical event occupies a
limited extension in space-time, in contrast with both the unextended 'elementary event',
which occupies only an infinitesimal portion of space and time, and with the event in
Whitehead's sense, which can fill an unlimited spatio-temporal region. The burning of a
body, the expansion of a gaseous mass, the refraction of a light ray, the boiling of a liquid,
the discharging of an electrical conductor, the fall of a meteor are examples of physical
events.
121 It is clear that, instead of this group of fundamental properties, we can choose a
different one, provided that the properties of the first group can be expressed in terms of
those of the second. Thus, instead of the intensities of the electromagnetic field, we could
consider the two electromagnetic potentials. We shall explain on pp. 201-02 what must be
understand by the proposition that one intrinsic property of one event is ex-pressed as a
function of other properties of this event. - It will be noticed that we have not listed the
distribution of material velocities (and that of the derivatives of field intensities) which it
exhibits, among the intrinsic properties of an event, even though this distribution might
appear to be decisive for its causal properties. But the explicit
knowledge of velocities and other derivatives seems inevitable to the physicist for deter-
mining the effects of a given distribution only because he operates on instantaneous
distributions: it is implicitly contained in the knowledge of 'a distribution exhibited by an
event extended in time.
122 Notice the rigorous reproduction of the refraction of a beam (or of any other physical
event) is highly improbable, and moreover, could not be experimentally deter-mined, even
if it should really occur. What makes the causal relation between these two
NOTES 295

optical events practically ascertainable is the continuity of this causal relation, which consists in
the fact that an approximate reproduction of the cause (i.e., the realization of an event which
exhibits a distribution of velocities and field intensities near to that of the cause) entails an
approximate reproduction of the effect. Therefore, it would be better to say that event A is a
deterministic cause of event B if every event A ' having in-trinsic properties sufficiently near
those of A is followed by an event B', whose intrinsic properties differ as little as we like from
those of B.
123 I.e., having no part in common withA.
124 We conclude from the principle of action by degrees as applied to time, that the past acts
upon the future only through the intermediary of the present, and that, conse-quently, a past part
of the cause can act upon the effect only through the effect of its
present part, provided that the latter contains all of the effects of the past part. This past portion
can therefore be removed from the cause without the latter's ceasing to deter-mine the effect.
Since each event is extended in time, and, consequently, can be decom-posed in an infinite
number of ways into two parts, one of which is past, when the other is present, we conclude that
there is in general no event which ceases to determine a given event upon the removal from it of
an arbitrarily chosen part.
125 Cf. S. Lesniewski, Podstawy ogolnej teorii mnogosci [Foundations of general set theory] I
Moscow, 1916.
126 The application of the causal relation to elementary events naturally presupposes an
appropriate passage to the limit. If, instead of considering the effect B of an event A,
we looked at a sequence B 1 , B2 , ... Bn ... of events whose spatio-temporal dimensions decrease
indefinitely and all of which have only one elementary event Bo in common, we
will establish that the corresponding sequence of causes A b A 2 , ... An ... also tends toward a
limiting event Ao. We will say of an arbitrary elementary event contained in
the limiting event Ao that it bears a causal relation to eventBo. It is clear that the same reasons
which guarantee the symmetry of the causal relation between extended events also entail the
symmetry of the causal relation between elementary events.
127 With regard to quantum events, cf. Part III below.
128 R. Ingarden, 'Von formalen Aufbau des individuellen Gegenstandes', sections 24-
25, Studia Philosophica, I (1935), 29-106.
129 Z. Zawirski, L 'evolution de la notion du temps. Cracow, Gebethner and Wolff,
1936, p. 207. (Referred to hereafter as Evolution.)
130 Cf. A. N. Whitehead,Enquiry, p. 75.
131 Cf. I, pp. 101-102.
132 Or also in terms of invariant simultaneity, cf. I, p. 95.
133 Cf. H. Minkowski, 'Raum und Zeit' in Lorentz, Einstein, Minkowski, Das Rela-
tivitiitsprinzip. 5th ed. Leipzig, Teubner, 1923, p. 55. (The Principle of Relativity; a
Collection of Original Memoirs on the Special and General Theory of Relativity, by H. A.
Lorentz, A. Einstein, H. Minkowski and H. Weyl, with notes by A. Sommerfeld, tr. by W. Perrett
and G. B. Jeffrey, London, Methuen, 1923.
134 The Events A, B, C are assumed to be distinct in this definition. By the effect of
an event X, we understand an arbitrary event causally related to X, i.e., a 'generalized effect' of
X (cf. I, p. 132 and Theorem 3, p. 215).
135 Cf.I,p.47.
136 Cf. Theorem 21, p. 235. By employing the notion of spatial coincidence, we can even
determine a decomposition in terms of a single pair of events (cf. Theorem 20, p. 234).
296 NOTES

137 Cf. Theorem 19, p. 234.


138 I, p. 131. Cf. also Pt. III.
139 Axiom II 3 is indispensable only on the assumption that the Euclidean nature of
momentary spaces referred to in II 4 is expressed in terms of betweeness and congruence.
It is well known, however, since the classic work of O. Veblen (,A System of Axioms for
Geometry', Trans. Am. Math. Soc. 5, 104.) that the entire Euclidean structure can be
concisely and elegantly axiomatized in terms of betweeness alone, if congruence is de-fined
projectively according to Cayley and Klein. Axiom II 3, which states that the mapping of two
instantaneous spaces upon one another obtained through spatial coin-cidence is isomorphic with
respect to spatial betweeness, would then suffice to guarantee that if geometry is Euclidean at a
single instant it will always remain so. Some of his axioms (e.g., 'I' and 'II') would be redundant
in formulating our II 4 because they follow from groups I and III. However, in evaluating this
alternative approach one has to con-sider that the projective definition of congruence seems very
remote from the scientist's actual mensurative procedures, in contradistinction to defmition D 6,
which leads directly to practical criteria,

140 Cf. Zawirski, Evolution, sections 27 -28.


141 Cf. I, pp. 61-62.
142 Cf. K. Ajdukiewicz, 'Das Weltbild und die Begriffsapparatur', Erkenntnis 4,
275.
143 To show the compatibility of axioms I-III we can use this arithmetical interpreta-tion
of the causal relation.
144 In other words, the necessary and sufficient condition for event B to be found
linearly between events A and C in the system where these are simultaneous, is that every
effect common to A and C be an effect of B. This statement can therefore be used as a
defInition of the linear order.
145 Proof. Assume that X is an effect of Yand that there is an event Z located between
X and Y (Le., acting upon every effect common to X and Y). If Z were not causally related to
X, there would be a decomposition in which they are simultaneous, by Axiom I 2. Since event Y
is outside of X and Z, it also would be simultaneous with X in this decomposition, by Defmition
D 3; there would therefore be two events which are simul-taneous and causally related, contrary
to the same defInition. We conclude that since X is causally related to Y, Z would also be
causally related to X. Similarly, we would con-
clude that Z is causally related to Y, Le., it is a common effect of X and Y. If Z acted upon
every effect common to X and Y, it would be its own effect, contrary to Theorem
2. It is therefore impossible that there should exist an event Z which acts upon every effect
common to X and Y, if the latter two are causally related.
146 The parameter of this transformation
tu-tv
v=
xU-xV
would then satisfy the inequality

Ivl< 1.

147 Cf. p. 234.


148 Cf. I, p. 140.
NOTES 297

149 It is sometimes added that the observer consults a clock attached to the trihedron. But it is
clear that the clock is not part of the reference system, no mqre than a grad-uated ruler used to
measure the spatial distances separating the event from the three coordinate axes. We omit the
latter by assuming that the axes themselves are graduated; but in the same way we can assume
that the temporal axis is graduated, i.e., that each event forming part of the history of the
trihedron is provided not only with its three spatial coordinates, but also with its temporal
coordinate. These two postulates taken together concern the system of coordinates and not the
reference system.
150 Cf. E. Mach, Die Mechanik im ihrer Entwicklung. 7th ed. Leipzig, Brockhaus, p. 232.
(The Science of Mechanics, tr. by T. J. McCormack. 6th ed., with revisions through the 9th
German ed. LaSalle, m., Open Court, 1960.)
151 H. Reicnenbach, Axiomatik der relativistichen Raum-Zeit Lehre [Die
Wissenschaft, 72]. Braunschweig, Vieweg, sections 10, 11 (Axiomatization of the Theory
of Relativity, tr. by Maria Reichenbach. Berkeley, University of California Press, 1969.)
152 Light can be defined in terms of the causal relation by supposing in conformity with the
structure of matter and electricity, that the elementary particles, e.g., the electron and proton, are
definable in geometrical terms, their diameters for example (which are, in fact, of different
orders of magnitude). Since the spatio-temporal distribution of the elec-trical charges is
expressed as a function of the causal relation, we will consider Maxwell's equations as defining
the components of the field vectors as a function of the distribu-tion of the charges. Poynting's
vector, formed with them, will in its turn define the light ray. Obviously, these def'mitions can
be critized for effecting experimental data very far removed from the direct data of experience
(e.g., the radius of the electron). However, these are traits inherent in the axiomatization of a
ready-made theory (cf. I, p. 100).
The axiomatic theory would nevertheless remain verifiable since the macrophysical
consequences arising from it are directly accessible to experience. The epistemological
interest of these definitions consists in the fact that they seem to prove that all physical
notions can be defmed as a function of the single causal relation.
153 Set t will form a complete instant of Il' since every event external to t has at least one effect
there. Therefore, t cannot be enlarged without introducing in it a pair of causally related events.

154 H. Weyl, Philosophie der Mathematik und Naturwissenschaft. Munich, Oldenbourg, 1928,
p. ISS. (Philosophy of Mathematics and Natural Science. Rev. and aug. Eng. ed., based on a
translation of Olaf Helmer. Princeton, Princeton University Press, 1949.
ISS Cf. for example, W. Wundt, Grundziige der physiologischen Psychologie, 3, Leipzig,
Engelmann, 1911, p. 260 ff. (Referred to hereafter as (Grundzfige.)
156 If one is prepared to accept psychophysical causality, he may, as Twardowski has shown,
define the expressive relation directly in terms of the causal relation - which fact further points
up the kinship between these two relations. cr. K. Twardowski, Rozprawy i artykuy
(Lectures and articles), Lwdw, Ksiegarnia SA 'Ksiaznica-Atlas' T. N. S. W., 1928, p. 114.

157 cr. Chapter 2, Section 2.


158 This is what is done in experimental psychology in the method of reactions. Cf. W.
Wundt, Grundzfige, 3, p. 359.
159 Cf. A. Sollberger, 'Biological Measurements in Time' in Interdisciplinary Perspectives
of Time. Annals of the New York Academy ofSciences 138, 2 (1967), 561-599.
160 Cf. section 4 of this Chapter.
298 NOTES

161 This is suggested by the circumstance that there are apparently homogeneous psychological
states capable of enduring without intrinsic change which correspond to periodic physical
processes: a sensation of the color red, for example, seems to be identical at each instant of its
duration, even though it does not correspond to an en-during physical state.

162 Wundt, Grundzuge, 3, p. 752.


163 . This condition is perhaps satisfied by sensory phenomena, of which no psychological
state seems to be totally free.
164 If ~ were not prior to r, it would be simultaneous with or posterior to it, according to
the hypotheses of the connectedness of succession and simultaneity. But it can-not be
simultaneous with it, since this would entail the simultaneity of r with 1/, nor posterior to
it, since this would imply the priority of r with respect to 1/, contrary to the hypothesis.
165 Cf. N. Kleitman: Sleep and Wakefulness. Chicago, University of Chicago Press, 1939;
rev. and enl. ed., 1963.
166 Thus, memory alone is the sleeper's best means of localizing his dream of the
night preceding awakening, while, by considering the substratum of this dream, we could
localize it with less than a second's error. Cf. Wundt, Grundziige. 3, p. 392. To realize the
advance due to the use of electroencephalograms, Cf. P. Verdone, 'Temporal Re-ference of
Manifest Dream Content', Perceptual and Motor Skills 20 (1965), 1253-
1268.
167 Cf. L. Busse, Geist und Korper, Seele und Leib. Leipzig, Diirr, 1903.
168 Kliiver, H. 'Mechanisms of Hallucinations' in Studies in Personality, contributed in
honor of Lewis Terman. New York, McGraw-Hill, 1942. 'Neurophysiology of Perception'
in Psychopathology of Perceptions; the Proceedings of the 53rd annual meeting of the
American Psychopathological Association, ed. by P. H. Hoch and J. Zubin. New York,
Grune and Stratton, 1965.
169 Of particular relevance are the following works by these authors:
J. C. Eccles, 'Brain and Free Will' in Consciousness and the Brain, ed. by G. G. Globus, G.
Maxwell and I. Savodnik. New York, Plenum Press, 1976.
D. J. Ingle and B. L. Baker, Physiological and Therapeutic Effects of Corticotrophin
(ACTH) and Cortisone. American Lecture Series, 179. Springfield, Ill., C. Thomas, 1953. N.
Kleitman, Sleep and Wakefulness. Chicago, University of Chicago Press, 1939; rev. and
enl. ed., 1963.
W. Penfield, 'Studies of the Cerebral Cortex of Man' in Brain Mechanisms and
Conscious-ness. 1954.
W. Penfield and L. Roberts, Speech and Brain Mechanisms. Princeton, Princeton
Univer-sity Press, 1959.
A. Rechtschaffen, 'Current Research on Sleep and Dreams'. 1966.
W. R. Sperry, 'Mental Phenomena as Causal Determinants in Brain Functions' in Con-
sciousness and the Brain, ed. by G. G. Globus, G. Maxwell and I. Savodnik. New York,
Plenum Press, 1976.
W. B. Webb and H. W. Agnew, 'Stage 4 Sleep', Science 174 (1971),1354-1356.
170 H. Feigl, The 'Mental' and the 'Physical'. Minneapolis, University of Minnesota Press;
1967.
171 H.von Foerster: 'Consciousness and Symbolic Processes' (4th Conference on Prob-lems of
Consciousness of the JosialI Macy Foundation), 1953.
NOTES 299

172H. Hoagland: 'Some Biochemical Considerations of Time,' in the Voices of Time,


ed. by J. T. Fraser, New York, Braziller, 1966, pp. 312-329.
173 J. von Neumann, Quantum Mechanics.
174 Mind-Energy, tr. by H. Wildon Carr. Westport, Conn., Greenwood Press, 1975 [reprint of
the New York, 1920 ed.), pp. 231-232.
INDEX TO VOLUME I

Aeson 85-86,290 Broglie, 1. de 6, 20


Agnew, H. W. 298 Brouwer, 1. E. J. 21
Ajdukiewicz, K. 226,296 Bunge, M. lO
Akhiezer, A. 1. 97 Busse, 1. 298
Ampere, A. M. 5
Aristotle 42,44 Campbell, N. R. 5
Arrhenius, S. 266 CaratModory, C. 8, 23, 25, l30, 189,
Aster, E. 286 192,193,294
Augustine 43 Carnap, R. 5, l3, 40, 42, 89, 95-100,
118, l37, 144-145,215,219,290,
Baker, B. 1. 298 291,294
Balmes, J. 70 Casella, R. C. 35
Bass, R. 291 Cassirer, E. 42,286, 287
Bateman, H. 23,189 Cauchy, A. 274
Baumann, J. 42,286 Cayley, A. 296
Benson, A. 290 Chandrasekhar, S. 33-35,151,152
Benussi, V. 293 Church, A. 21
Berestetskii, V. B. 97 Chwistek, 1. 29
Bergmann, H. 132-133, 286, 290, 292 Clarke, S. 42,43, 286
Bergson, H. 5, 13, 43, 86, U8, 123, Clausius, R. 31,83,130,148
168,219-220,224,239,259,262- Cohen, J. 293
263, 265, 271-272, 275, 289,290, Cohen, R. S. 286,291
291 Conforto, G. 35
Berkeley, G. 89 Costa de Beauregard, O. lO, 147, 153,
Bernays, P. 21,29 154-155
Bernoulli, D. 21 Couturat, L. 287
Bernoulli, J. 21 Cunningham, E. 23
Biema, E. van 286,287
Birkhoff, G. D. 21 Darwin, C. 25,241
Bohm, D. 22 De Morgan, A. 9
Bohr, N. 5,6,17,20,22 Descartes, R. 43,45, 280
Boltzmann, 1. 25,129,130,148 de Sitter, A. 152
Bondi, H. 154 Destouches-Fevrier, P. See Fevrier, P.
Born, M. 137 Dirac, P. A. M. 6, 8, 14, 159, 193,282,
Bosanquet, B. 288 294
Bose, S. N. 13,290 Drews, A. 287
Bourguet,1. 45 Dumas, G. 276
Brentano, F. 168 Dyson, F. J. 6
Bridgman, P. W. 5,9,13,292
Broad, C. D. 23,28,289 Eccles,J.C.13,262,264,298

300
INDEX TO VOLUME I 301
Eddington, A. S. 5, 25,26,31, 153, Haag, R. 8

158,196,292 Haldane, J. B. S. 25
Ehrenfest, P. 25,154,156 Hamilton, W. R. 131
Ehrenfest, T. 25, 156 Heidegger, M. 290
Eibl, H. 286 Heisenberg, W. 6,17,20,129,154
Einstein, A. 3,4,8,9,12,13,16,17, Hertz, P. 294
22, 23, 26, 32, 33, 34, 40, 86, 91, Heyting, A. 21
96, 105, 108, 109, 110, 123, 124, Hilbert, D. 21,29,33,102-103
125, 145, 148, 149, 150, 151, 152, Hoagland, H. 13, 179-180, 265 -266,
153, 154, 158, 189, 213, 215, 216, 293, 299
234,236,291,295 Hoch, P. H. 298
Enriques, F. 224,289 Hoyle, F. 154
Espagnat, B. d' 10 Hume, D. 39,57,79, 116, 117, 119,
Euclid 91 197,278,291
Husser!, E. 56, 290, 293
Faraday, M. 117,206
Fechner, G. 261 Ignatovsky, W. von 23
Feigl, H. 258,265,298 Infeld, L. 189,190
Feinberg, G. 35 Ingarden, R. 205,295
Fermi, E. 13, 290 Ingle, D. J. 13,258,262,265,266,284,
Fevrier, P. 10,21 298
Feyerabend, P. K. 10
Feynman, R. P. 6,8,32, 158 Janet, P. 293
Fisher, R. A. 25 Jason 290
Fock, V. A. 23,33 Jauch, J. M. 21
Foerster, H. von 13, 179-180, 265, Joan of Arc 163
266, 298 Jordan, P. 76-77,103,111,117-118,
Fraisse, P. 293 119, 120
Frank, P. 23
Fraser, J. T. 157,293,299 Kant, I. 39,40, 41, 42, 43, 44, 48,49,
Frege, G. 21,28 50, 51-60,62-68,70,73-74,75,
Fresnel, A. 20 88, 89-90, 94, 126, 143, 145, 194,
205,219,280,287,288,291
Gel'fand, I. M. 11 Kelvin, W. T. 5
Gell-Mann, M. 6 Khinchin, A. I. 25, 156
Gerhardt, C. I. 286 Klein, F. 296
Gibbs, J. W. 25,129 Kleitman, N. 13, 262, 265, 266, 293,
Globus, G. G. 298 298
Godel, K. 8, 10, 29, 33-35, 148-151, Kltiver, H. 13, 262, 298
152,153,264 Kolmogorov, A. N. 21, 159
Gold, T. 34, 154 Kotarbinski, T. 289
Goldstone, S. 293
Gombaud, Antoine, Chevalier de Mere Lagrange, J. L. 148, 155, 156, 220
21 18-19 Lange, F. 232
Goodman, N. Lapicque, L. 276
Grtinbaum, A. 10, 147, 153, 154- Larmor, J. 91
155 Layzer, D. 160
302 INDEX TO VOLUME I

Lechalas, G. 48, 65, 70-71, 73-80, 83-84, Newton, I. 4,8,11,25,43,52,66,67, 270


86, 88-90, 95, 103, Ill, 113, 126, 129,
130, 131, 145,147, 288,289,290,292
Occam See William of Occam
Lehmann, H. E. 8,293 Ovid 290
Leibniz, G. 40,41,42-46,48-50,51, 64, 68, Ozav:ith, L. 151
70, 80, 96, 117, 134, 135, 145, 188, 213,
219, 253, 261, 265, Pascal, B. 21
271,280,286,287,288 Pauli, W. 16,17,22,23
Pauling, L. 5
Lesniewski, S. 201,295 Penfield, W. G. 13,258,259,262-263,
Lewin, K. 40,42,83,291 265,266,284,285,298
Lewis, G. N. 137,155 Pertz, G. H. 42
Locke, J. 89 Perutz, M. 5
Loemker, L. E. 286 Pillon, F. T. 288
Lorentz, H. A. 91,192,295 Planck, M. 1,9,22
Lotze, H. 70 Plato 5,84
Liiders, G. 16 Poincare, H. 5,74,143,224,226
Lutoslawski, W. 5 Poly a, G. 32
Popper, K. R. 5, 154, 159
Mach, E. 26,118,297 Poynting, J. 297
Mackey, G. W. 8
McTaggart, J. E. 149 Quine, W. V. o. 18,29
Malebranche, N. 79
Margenau, H. 10, 13 Ramsey, F. P. 29
Marie, P. 262 Rechtschaffen, A. 13, 258, 262, 264,
Maupertuis, P. 8 265,266,284,293,298
Maxwell, G. 298
Maxwell, J. 8, 20, 23, 25, 117, 129, Reichenbach, H. 5, 8, 13, 21, 23, 40, 42,
141,297 64, 65, 83, 95, 101, 105-108, 110-113,
Medea 85,290 115-116, 118-ll9, 121, 123-129, 132,
Meinong, A. 291 133, 137,144-145, 147, 154, 189, 190,
Mere, Chevalier de See Gombaud, 192,215,219, 232,286,289,291,297
Antoine, Chevalier de Mere
Meyerson, E. 5, 13 Reichenbach, M. 286,291,297
Michelson, A. A. 17 Renouvier, C. B. 288 Rickert, H.
Minkowski, H. 16,23,27,29,91,93, 32
107,116,149,210,295 Robb, A. A. 23,40,48,91-100, ll9, 124, 137,
Mises, R. von 158 144, 145, 215, 219, 290, 291
More, H. 43
Mozart, W. A. 178 Roberts, L. 298
Roosevelt, F. D. 13
Nagel, E. 10 Rosenfeld, L. 17
Napoleon 135,136 Rothe, H. 23
Neeman, Y. 35 Ruelle, D. 8
Neumann, J. von 5, 8, 10, 13, 15, 20, Russell, B. 5, 10, 13, 21, 27, 28, 29,
21,27,33,129,156,159,193,258, 266- 53,56,59,80,83,93,98,106,120, 121, 126,
267,285,294,299 128, l31, l34-145, 258,
INDEX TO VOLUME I 303
288,289,290,291,292,293 Vaihinger, H. 193

Rutherford, E. 4 Varignon, P. 287


Veblen, O. 296
Sachs, R. G. 31,35,155 Verdone, P. 298
Savodnik,1. 298 Vigier, J.-P. 22
Schaltenbrand, G. 293 Vries, H. de 25, 156, 179
Schilpp, P. 148
Schlegel, R. 155-157 Wald, A. 14, 159
Schlick, M. 13 Watanabe, S. 155-156,157-160
Schopenhauer, A. 39, 54, 58, 59-65, Watson, J. 5
94,97,205,242,287 Webb, W. B. 262,298
Schrodinger, E. 6,15,34,129,154 Weber, W. 261
Schwinger, J. S. 6,16, 147 Weizsiicker, C. F. von 17
Shapere, D. 10 Weyl, H. 23, 26, 33, 154, 189, 192,
Sherrington, C. 276 236,294,295,297
Smoluchowski, M. von 25,86,130,154, Whitehead, A. N. 13, 21, 27-28, 126,
156,292 128, 135, 175, 201, 206,207-208,
Sollberger, A. 297 209,290,291,293,295
Sperry, W. R. 262-263,264,298 Whitrow, G. 1. 147,152-153
Spinoza, B. 43,249 Wiener, N. 10,29
Stein,H. 34 Wightman, A. S. 8
Steinhaus, Z. 285 Wigner, E. 5,6,10,15,155, 156
Stevens, S. 14 Wilde, O. 259
Stiegler, W. 23 William of Occam 53
Stroud, J. M. 293 Wittgenstein, L. 27
Suppes, P. 8 155-157 Wolff, c. 42 33-34
Swinburne, R. Wright, J. P.
Symanzik, K. 8 Wundt, W. 253,297,298
Szilard, L. 13
Yukawa, H. 6
Taine, H. 14
Tarski, A. 2,10,28,29,264,291 Zawirski, A. 206,295,296
Teller, E. 13 Zeman, J. J. 21
Tomonaga, S. 6 Zermelo, E. 25
Twardowski, K. 297 Zilsel, E. 290
Zimmermann, W. 8
Urey, H. C. 24 Zubin, J. 298
BOSTON STUDIES IN THE PHILOSOPHY OF SCIENCE

Editors:
ROBERT S. COHEN and MARX W. WARTOFSKY
(Boston University)

I.Marx W. Wartofsky (ed.), Proceedings of the Boston Colloquium for the


Philosophy of Science 1961-1962.1963.
2. Robert S. Cohen and Marx W. Wartofsky (eds.), In Honor of Philipp Frank. 1965.
3.Robert S. Cohen and Marx W. Wartofsky (eds.), Proceedings of the Boston Collo-
quium for the Philosophy of Science 1964-1966. In Memory of Norwood
Russell Hanson. 1967.
4. Robert S. Cohen and Marx W. Wartofsky (eds.), Proceedings of the Boston Collo·
quium for the Philosophy of Science 1966-1968. 1969.
5. Robert S. Cohen and Marx W. Wartofsky (eds.), Proceedings of the Boston Collo-
quium for the Philosophy of Science 1966-1968. 1969.
6. Robert S. Cohen and Raymond J. Seeger (eds.), Ernst Mach: Physicist and Philos-
opher. 1970.
7. Milic Capek, Bergson and Modern Physics. 1971.
8. Roger C. Buck and Robert S. Cohen (eds.), PSA 1970. In Memory of Rudolf Carnap.
1971.
9. A. A. Zinov'ev, Foundations of the Logical Theory of Scientific Knowledge
(Complex Logic). (Revised and enlarged English edition with an appendix by G. A.
Smirnov, E. A. Sidorenka, A. M. Fedina, and L. A. Bobrova.) 1973.
10. Ladislav Tondl, Scientific Procedures. 1973.
11. R. J. Seeger and Robert S. Cohen (eds.), Philosophical Foundations ofScience. 1974.
12.Adolf Griinbaum, Philosophical Problems of Space and Time. (Second, enlarged
edition.) 1973.
13.Robert S. Cohen and Marx W. Wartofsky (eds.), Logical and Epistemological
Studies in Contemporary Physics. 1973.
14. Robert S. Cohen and Marx W. Wartofsky (eds.), Methodological and Historical
Essays in the Natural and Social Sciences. Proceedings of the Boston
Colloquium for the Philosophy of Science 1969-1972. 1974.
15. Robert S. Cohen, J. J. Stachel and Marx W. Wartofsky (eds.), For Dirk Struik.
Scientific, Historical and Political Essays in Honor of Dirk Struik. 1974.
16. Norman Geschwind, Selected Papers on Language and the Brain. 1974.
18. Peter Mittelstaedt, Philosophical Problems of Modern Physics. 1976.
19. Henry Mehlberg, Time, Causality, and the Quantum Theory (2 vols.). 1980.
20.Kenneth F. Schaffner and Robert S. Cohen (eds.), Proceedings of the 1972 Biennial
Meeting, Philosophy of Science Association. 1974.
21. R. S. Cohen and J. J. Stachel (eds.), Selected Papers of Leon Rosenfeld. 1978.
22. Milic Capek (ed.), The Concepts of Space and Time. Their Structure and Their
Development. 1976.
23. Marjorie Grene, The Understanding of Nature. Essays in the Philosophy of Biology.
1974.
24. Don Ihde, Technics and Praxis. A Philosophy of Technology. 1978.
25.Jaakko Hintikka and Unto Remes, The Method of Analysis. Its Geometrical Origin
and Its General Significance. 1974.
26. John Emery Murdoch and Edith Dudley Sylla, The Cultural Context of Medieval
Learning. 1975.
27. Marjorie Grene and Everett Mendelsohn (eds.), Topics in the Philosophy of Biology.
1976.
28.Joseph Agassi, Science in F7ux. 1975.
29. Jerzy 1. Wiatr (ed.), Polish Essays in the Methodology of the Social Sciences. 1979.
32. R. S. Cohen, C. A. Hooker, A. C. Michalos, and J. W. van Evra (eds.), PSA 1974: Proceedings
of the 1974 Biennial Meeting of the Philosophy of Science Association.
1976.
33. Gerald Holton and William Blanpied (eds.), Science and Its Public: The
Changing Relationship. 1976.
34.Mirko D. Grmek (cd.), On Scientific Discovery. 1980.
35. Stefan Amsterdamski, Between Experience and Metaphysics. Philosophical
Problems of the Evolu tion of Science. 1975.
36. Mihailo Markovic and Gajo Petrovic (eds.), Praxis. Yugoslav Essays in the
Philosophy and Methodology of the Social Sciences. 1979.
37. Hermann von Helmholtz: Epistemological Writings. The Paul Hertz/Moritz Schlick
Centenary Edition of 1921 with Notes and Commentary by the Editors. (Newly
translated by Malcolm F. Lowe. Edited, with an Introduction and Bibliography, by Robert S.
Cohen and Yehuda E1kana.) 1977.
38. R. M. Martin, Pragmatics, Truth, and Language. 1979.
39. R. S. Cohen, P. K. Feyerabend, and M. W. Wartofsky (eds.), Essays in Memory of Imre
Lakatos. 1976..
42. Humberto R. Maturana and Francisco 1. Varela, Autopoiesis and Cognition. The
Realization of the Living. 1980.
43.A. Kasher (ed.), Language in Focus: Foundations, Methods and Systems. Essays
Dedicated to Yehoshua Bar·Hillel. 1976.
46. Peter L. Kapitza, Experiment, Theory, Practice. 1980.
47. Maria L. Dalla Chiara (ed.), Italian Studies in the Philosophy of Science. 1980.
48. Marx W. Wartofsky, Models: Representation and the Scientific Understanding.
1979.
50. Yehuda Fried and Joseph Agassi, Paranoia: A Study in Diagnosis. 1976.
51. Kurt H. Wolff, Surrender and Catch: Experience and Inquiry Today. 1976.
52. Karel Kosik, Dialectics of the Concrete. 1976.
53. Nelson Goodman, The Structure of Appearance. (Third edition.) 1977.
54. Herbert A. Simon, Models ofDiscovery and Other Topics in the Methods of Science.
1977.
55. Morris Lazerowitz, The Language of Philosophy. Freud and Wittgenstein. 1977.
56. Thomas Nickles (ed.), Scientific Discovery, Logic, and Rationality. 1980.
57. Joseph Margolis, Persons and Minds. The Prospects of Nonreductive Materialism.
1977.
58. Gerard Radnitzky and Gunnar Andersson (cds.), Progress and Rationality in Science.
1978.
59. Gerard Radnitlky and Gunnar Andersson (cds.), The Structure and Development of
Science. 1979.
60. Thomas Nickles (ed.), Scientific Discovery: Case Studies. 1980.
61. Maurice A. Finocchiaro, Calileo and the Art of ReasoninK. 1980.

Você também pode gostar