You are on page 1of 262

PHYSICAL SCIENCES AND HISTORY OF PHYSICS

BOSTON STUDIES IN THE PHILOSOPHY OF SCIENCE

EDITED BY ROBERT S. COHEN AND MARX W. WARTOFSKY

VOLUME 82
PHYSICAL SCIENCES
AND
HISTORY OF PHYSICS
Edited by

R. S. COHEN and M. W. WARTOFSKY

D. REIDEL PUBLISHING COMPANY


~
A MEMBER OF THE KLUWER " ACADEMIC PUBLISHERS GROUP

DORDRECHT/BOSTON/LANCASTER
Library of Congress Cataloging in Publication Data

Main entry under title:

Physical sciences and history of physics.

(Boston studies in the philosophy of science; v. 82)


Bibliography: p.
Includes index.
1. Physics-History. I. Cohen, Robert Sonne.
II. Wartofsky, Marx W. III. Series.
QC7.P46 1983 530'.09 83-21352
ISBN-I3: 978-94-009-7180-6 e-ISBN-I3: 978-94-009-7178-3
001: 10.1007/978-94-009-7178-3

Published by D. Reidel Publishing Company,


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

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


by Kluwer Academic Publishers,
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.

All Rights Reserved.


© 1984 by D. Reidel Publishing Company
and copyright holders 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 information storage and
retrieval system, without written permission from the copyright owner.
EDITORIAL PREFACE

These essays on the conceptual understanding of modern physics strike


directly at some of the principal difficulties faced by contemporary philos-
ophers of physical science. Moreover, they reverberate to earlier and classical
struggles with those difficulties. Each of these essays may be seen as both a
commentary on our predecessors and an original analytic interpretation.
They come from work of the past decade, most from meetings of the Boston
Colloquium for the Philosophy of Science, and they demonstrate again how
problematic the fundamentals of our understanding of nature still are. The
themes will seem to be familiar but the variations are not only ingenious but
also stimulating, in some ways counterpoint. And so once again we are
confronted with issues of space and time, irreversibility and measurement,
matter and process, hypothetical reality and verifiability, explanation and
reduction, phenomenal base and sophisticated theory, unified science and
the unity of nature, and the limits of conventionalism.
We are grateful for the cooperation of our contributors, and in particular
for the agreement of George Ellis and C. F. von Weizsiicker to allow us to
use previously published papers.

October 1983

Boston University R. S. COHEN

Baruch College, City University of New York M.W.WARTOFSKY


T ABLE OF CONTENTS

EDITORIAL PREFACE v
ACKNOWLEDGEMENTS ix

M ILIC CAPEK / Particles or Events?


PATRICK A. HEELAN / Commentary on 'Particles or Events?' 29
O. COSTA DE BEAUREGARD / Time Symmetry and Interpreta-
tion of Quantum Mechanics 35
HANS EKSTEIN / Is Physical Space Unique or Optional? 57
CATHERINE Z. ELGIN / Theory Reduction: A Question of Fact or
or a Question of Value? 75
GEORGE F. R. ELLIS / Cosmology and Verifiability 93
DAVID HEMMENDINGER / Galileo and the Phenomena: On Mak-
ing the Evidence Visible 115
GEN-ICHlRO NAGASAKA / Quantum Theory of Measurement:
A Non-Quantum Mechanical Approach 145
JOACHIM PFARR / Protophysics of Time and the Principle of
Relativity 159
PETER JANICH / Commentary on 'Proto physics of Time and the
Principle of Relativity' 191
PAUL M. QUAY / Temporality and the Structure of PhYSics as
Human Endeavor 199
S. S. SCHW EB E R / Commentary on 'Temporality and the Struc-
ture of Physics as Human Endeavor' 231
C. F. v. WEIZSACKER / The Unity of Nature 239
INDEX OF NAMES 255
ACKNOWLEDGEMENTS

Two of the papers in this volume have appeared in earlier publications.


Permission to reprint them here is gratefully acknowledged.
The first is 'Time Symmetry and Interpretation of Quantum Mechanics'
by O. Costa de Beauregard, which was first published in the journal Founda-
tions of Physics, Vol. 6, No.5, pp. 539-559 (New York: Plenum Publishing
Corporation, 1976).
The second is 'Cosmology and Verifiability' by George F. R. Ellis, from
the Quarterly Journal of the Royal Astronomical Society, Vol. 16, No.3,
pp. 245-264 (Oxford and London: Blackwell Scientific Publications, 1975).

ix
MILlC CAPEK

PARTICLES OR EVENTS?

I believe I should start with a kind of opening statement which will make the
purpose of this paper clear and its presentation easier to follow. In the first
place, it is not going to be a paper on philosophy or methodology of science
- at least not in its usual, orthodox sense - but rather a philosophical com-
ment on one particularly significant trend in twentieth-century physics. You
may call it an essay in 'philosophy of nature', if we understand the term prop-
erly. I am fully aware of how unpopular and discredited this term has become;
it is now rare to fmd institutions which still offer courses in 'philosophy of
nature'. It really takes courage to do so and I commend my colleague Robert
Cohen for having introduced courses of this kind in the Boston University
curriculum. It is not difficult to trace the causes of this unpopularity and I
have analyzed them in some of my previous writings. In the first place, the
term itself is a translation of the German Naturphilosophie coined by the
German idealists in the post-Kantian period, and a lingering disappointment
with their speculative and arbitrary constructions comes immediately to mind
as soon as the word is mentioned. In truth, we could hardly find another
period in which the contrast between sterile and a priori speculations such as
those of Schelling and Hegel and the genuine progress in the empirical sciences
were more striking; we have only to consider the development of geology,
biology, chemistry and of the physics of electricity and magnetism during the
same period. Second, even if we understand 'philosophy of nature' in a more
acceptable and iess pretentious way as an attempt to synthesize various
scientific fields, that is, as 'completely unified knowledge' in the sense that
Herbert Spencer in the second half of the last century defined philosophy in
general, some grave doubts remain. When, after all, would scientific know-
ledge be fully unified? Spencer's name itself reminds us of how premature
and ambitious his attempt at a 'complete integration of knowledge' was; all
he achieved was a codification and integration of nineteenth-century scien-
tific knowledge and only in a rough and approximate sense. Isn't the same
thing likely to happen to anybody who would try to synthesize the scientific
knowledge so enormously increased and diversified, would not such an
attempt be even more unrealistic and more pretentious now?
Thus it is natural to restrict the philosophy of science to a mere method-
1
R. S. Cohen and M. W. Wartofsky (eds.), Physical Sciences and History of Physics, 1-28.
© 1984 by D. Reidel Publishing Company.
2 MILIC CAPEK

ology and to veto all questions having an even remotely metaphysical ring,
such as the questions raised by philosophy of nature undoubtedly have.
About fifty years ago, Moritz Schlick still had the courage to name one of
his books Philosophie der Natur, despite the fact that he belonged to the
generally anti-metaphysically oriented Vienna Circle. Today, even terms such
as 'the nature of the universe' bring a contemptuous smile to the lips of some
scientists and the majority of philosophers of science.
Nevertheless I still believe that philosophy of nature is a legitimate enter-
prise provided we are careful to redefme its task and its limits. There is no
question that there are certain definite trends in the sciences, for example, in
present-day physics, which remain either undiscerned or ignored when one
confines oneself to a certain narrow field of specialization or when interest is
restricted to questions of methodology only. Such trends can be discerned
only within a wider context - more specifically, when a broader historical
perspective is adopted. How can any trend be discerned without considering
its contrasting historical backdrop? How can the inadequacy of the concept
'particle', about which I want to speak, be discerned without first bringing
into focus as sharply as possible all its essential features and their relations to
other classical concepts? It may be argued that no trend can be established
beyond doubt since there have been many so-called 'trends' which proved to
be reversible, in the sense that they were eventually replaced by trends in an
opposite direction. I suppose that today when claims are made that there is
no progress being made at all in scientific knowledge, such a view is probably
very fashionable.
Such platitudinous generalities about the alleged reversibility of any
trend are possible only when all the evidence for the persistence of some
trends is disregarded - and such evidence is indeed massive. Furthermore,
even if we still regard such evidence as circumstantial, it is greatly streng-
thened by epistemological considerations which can hardly be ignored. In the
context of the problem I am going to discuss, it is not enough to show all the
growing evidence for the inadequacy of the concept 'particle' on the micro-
physical level; one must also show how the psychological origin of this con-
cept, which developed under the pressure of limited macroscopic experience,
makes its applicability to the 'microcosmos' exceedingly improbable.

I. THE FRUITFULNESS OF CORPUSCULAR EXPLANATIONS


IN CLASSICAL PHYSICS

The inadequacy of the concept of particle, or, at least its inapplicability


beyond the limits of our macroscopic experience had been suspected long
PARTICLES OR EVENTS? 3

ago, although mainly on epistemological, logical or metaphysical grounds. To


mention just two outstanding examples: l..eibniz in the eighteenth and Mach
in the nineteenth century, both philosopher-scientists, were severe critics of
atomism; one may say that the whole of their thought was pervaded, and in a
sense inspired, by a deep distrust of the concept of particle on which the
corpuscular-kinetic models of nature and, more generally, the whole mechanis-
tic view of the universe were based. But the fruitfulness and success of these
models in explaining various concrete phenomena were such that this criti-
cism was largely ignored, at least by physicists. In fact, even the .critics them-
selves could not remain blind to the triumphs of the mechanistic, i.e., the
corpuscular-kinetic models. To use the same examples, l..eibniz, while rejecting
the concept of atom (i.e., of an indivisible material particle) in his meta-
physics, at least favored mechanistic Cartesian models in his physics, and he
even accused Newton of smuggling an occult quality under the name of
'attracting force' back into nature. Matter was for him a "well-founded
phenomenon" (phaenomenon bene fundatum) which, while being constituted
of immaterial, non-mechanical entities (monads), nevertheless appears to us in
a way which can be described by mechanistic models. The case of Ernst Mach,
one of the severest critics of what he called "mechanistic mythology", is even
more instructive. He, followed in this respect by Ostwald, did his best to sub-
stitute more abstract energetic explanations for corpuscular models; but at
the end of their lives, neither of these men could ignore the overwhelming
empirical evidence in favour of the reality of atoms. In retrospect the whole
revolution against atomism and mechanism at the end of the last century,
characterized by the names Mach, Stallo, Ostwald and Duhem, was largely
premature, since contrary to the views of these men, the fruitfulness of cor-
puscular-kinetic explanations at that time had not yet been exhausted. Only
with the coming of the relativity theory and, even more conspicuously, with
the discoveries of the wave nature of matter, did the basic inadequacy of the
concept of particle become obvious and undeniable.
How serious the present crisis surrounding this concept is, will become
clearer when we realize how spectacular the previous success of the cor-
puscular-kinetic explanations of various phenomena had been. My time for
even a short historical digression is severely limited; but I would like to review
at least briefly the main phases of the development of atomism and what
appeared as its ultimate triumph.
It is generally known that the atomistic tradition, i.e., the view that matter
consists of ultimate, indivisible particles, can be traced to l..eucippus and
Democritus in the fifth century B.C.; it is less generally known that it had not
4 MILIC CAPEK

entirely disappeared even in the Middle Ages, though it was driven under-
ground by the pressure of the medieval establishment which accepted Aristo-
telian physics and cosmology; fmally, it re-emerged victorious during the
cosmological revolution of the sixteenth and seventeenth centuries and its
revival coincided with the foundation of modem classical science_ The con-
tinuity between Democritus, Gassendi, Dalton and even Lorentz is obvious to
any unprejudiced person who is acquainted with the history of ideas and is
doubted only by those who are not. (I am thinking in particular of such
historians of ideas as Cyril Bailey, Kurt Lasswitz, Emile Meyerson and
Federigo Enriques. 1 It is usually claimed, in particular by some rank-and-file
physicists, that there are essential differences between the ancient atomism
which was allegedly purely speculative and the modem one which is based on
extensive experimental verifications. 2 This objection, plausible as it may
appear, overlooks the fact that Greek atomism was born out of the reaction
against the metaphysics of Parmenides and Zeno of Elea, the metaphysics
which Benjamin Farrington appropriately characterized as a "reaction against
experiential science".3 In rejecting this reaction, the atomists returned to
concrete sensory experience, though not without retaining a large portion of
the conceptual apparatus of the Eleatic school. I believe it was Windelband
who said picturesquely that "Democritus smashed the Parmenidean sphere of
Being into tiny fragments and scattered them through empty space". In other
words, the atom of Democritus - and this remained true of the atoms of all
the periods up to the end of the nineteenth century - retained all the attri-
butes of the Eleatic Being: it was immutable, that is, indestructible, uncrea-
table, and indivisible; each atom was 'one' in the sense that it filled the
volume it occupied fully and continuously, in a homogeneous and undifferen-
tiated way. There was, of course, one fundamental difference: the atomists
recognized the reality of the void between the atoms in order to account for
our undeniable experience of diversity and change, - the experience which
Eleatic metaphysics was unable to explain and which it simply and arrogantly
denied.
It is precisely this feature which made classical atomism implicitly modem,
in the sense that it made its future impressive empirical verification possible.
There were two basic themes common to atomism in all its forms and phases
which have not changed through the centuries: that all diversity of nature is
reducible to the differences in configuration of the basic homogeneous
particles; and, second, that all apparently qualitative changes are reducible
to various displacements of the same basic units. In this respect classical
atomism was far superior to the physics of Aristotle, which upheld real quali-
PARTICLES OR EVENTS? 5

tative differences between the four - or rather five - heterogeneous elements


and which, in regarding, for instance, the process of evaporation as a real trans-
formation of one element - water - into another - air - also believed in real
qualitative changes in nature. The modem concept of matter as a homo-
geneous stuff differentiated only by the quantitative differences between the
ultimate particles - that is, their size, shape, position and motions - was
fully anticipated by Democritus while it remained completely foreign to
Aristotle's qualitative physics. There was a similar contrast between the
infinite space of the atomists, homogeneous in all its parts and isotropic, and
the naive spherical universe of Aristotle, differentiated into the celestial and
sublunar realms and into the heterogeneous concentric zones of "natural
places". There is no question now which of these views was superior and
which was closer to the spirit and even to the letter of Newtonian physics.
It is true that it took a considerable time before a more correct view pre-
vailed. As I mentioned, the atomistic view of matter was almost eliminated in
the Middle Ages when it was regarded as synonymous with atheism; it
survived outside of, and in opposition to, the medieval establishment which
made Aristotle's cosmology its own. No wonder the atomists were at that
time persecuted or silenced. Thus in 1348 Nicolas d' Autrecourt was forced by
order of the University of Paris to recant the following 'errors': that in nature
there is no coming into being nor any real annihilation, but merely changes in
position; that light consists in local motion which is propagated at a finite
velocity; that in nature there is no real generation nor any real destruction,
but merely congregation and disaggregation of atoms, etc. Note that all these
allegedly erroneous opinions proved to be correct - but only in the century
of Newton and Robert Boyle,4 that is, only after Lucretius's view of infinite
space was rehabilitated by Giordano Bruno, and atomism in general by Pierre
Gassendi and others. This started the process Dijksterhuis called "mechaniza-
tion of the world picture" which continued triumphantly and practically
without interruption until the end of the last century.
Thus the main difference between classical Greek atomism and modem
atomism was one of degree; the former was based on far more limited ex-
perience than the latter, that is, on the sensory experience of the naked eye
(and naked touch) unaided by telescope, microscope and other devices by
which the field of our perception has been enormously extended. Yet this
limited experience was analyzed by the early atomists so attentively and with
such a finesse that their main conclusions about the nature of the physical
world were identical to the conclusions of modem atomists. Anybody reading
Lucretius attentively and not merely for aesthetic or philological pleasure is
6 MILlC CAPEK

struck by the wealth of empirical facts he dealt with and by the acuteness
with which he analyzed them. Only inattentive and superficial readers with a
limited knowledge of the physical sciences can still maintain the fiction tl,lat
ancient atomism was "purely speculative". In truth, some ofits anticipations
were remarkably precise and astonishingly sp~cific: such as Lucretius's view of
hidden molecular motion, imperceptible to our senses, eighteen centuries
before its actual discovery by Brown and its correct interpretation by
Ramsey,S or Democritus' view that empirical differences in macroscopic
bodies are due to the differences in shape (axiiIlOl), arrangement (T(~~L~), and
position ((J€(Jt~) of the atoms 6 - this represents the same general approach to
the observed diversity of matter as that adopted by modem structural chemis-
try in its explanation of isomerism, polymerism, polymorphism; in its inter-
pretation of the diversity of chemical elements in terms of different electro-
nic configurations, etc.
It may be objected that such anticipations are too general to be significant,
or even to be called genuine anticipations. It is true that they were conse-
quences of certain general principles which both ancient and modem atomism
share. But the list of these general principles shows how important they are:
the homogeneity and infinity of space; the conservation and unity of matter;
the reduction of all empirical diversity to the differences in configuration and
motion; the reduction of all changes to motion. Add to this the fact that
ancient atomists came remarkably close to stating the laws of inertia and of
the conservation of momentum when they asserted that the motion of atoms
is as eternal as the atoms themselves and that each of them continues moving
along a straight path until it rebounds from other atoms. 7 If this may appear
vague and unsatisfactory to the modem mind, let us compare it to the
opposite view of Aristotle according to which every motion requires a mover,
i.e. a moving force to keep a body moving - a proposition which vitiated his
whole physics and a large part of his metaphYSiCS.
What happened during the period 1600-1900 is too well known to be
dwelt upon: there occurred a gradual but decisive empirical verification of all
important general insights of classical atomism: the homogeneity and infinity
of space, the unity and constancy of matter, the mechanization of the world
picture with all its corollaries. All continuous fluids such as phlOgiston,
caloric and the electric fluid disappeared from physics and even the last one
which remained - aether - was sometimes interpreted in a corpuscular, and
always in a mechanistic fashion. s By the end oflast century, the mechaniza-
tion of the world picture seemed nearly complete; even the electromagnetic
theory of matter which at first seemed to be a rival of the mechanistic view,
PARTICLES OR EVENTS? 7

proved to be its ally with the coming of electron theory. For even electrons
and protons were, after all, particles and although their motions were ruled
by the laws of mechanics combined with the laws of electromagnetism, the
hope that the latter laws could be interpreted mechanically by some appro-
priate model of the electromagnetic aether persisted until the advent of
relativity and quantum theory. But in its initial phase, electron theory ful-
filled one of the most cherished dreams of the atomistic-kinetic view - the
reduction of all qualitative differences to those of configuration: it explained
the diversity of chemical elements by the differences in the number of
nuclear particles and the corresponding number of orbital electrons. The
Democritean "alphabet of Being" seemed finally to be within our grasp.

II. THE CRISIS FOR CORPUSCULAR MODELS

Having said all this, I hope I am beyond any possible reproach of being unfair
to atomism or, more generally, to the corpuscular-kinetic view of nature. Yet,
it is precisely the very applicability of this view outside our macroscopic
experience which is now in doubt; and it is in doubt because the concept of
an indivisible, permanent corpuscle moving along a continuous trajectory
through space and identifiable at successive instants in time seems to be
utterly inadequate on the microphysical scale. At least the circumstantial
evidence pointing in this direction is overwhelming and still increasing.
For we all know what happened after 1900 to that impressive looking
edifice of classical physics which had appeared nearly complete: not only did
its roof collapse, but even its very foundations had to be rebuilt. All funda-
mental classical concepts had to be either given up or profoundly revised:
space, time, matter, motion, causality. It is very difficult to deal separately
with each of these concepts since a change of one is related to the changes in
the others. This follows from the fact that all these concepts were related in a
particular way and in most instances their very defmitions contained
references to other concepts. What we called the corpuscular-kinetic model of
nature had been a precise, conceptual network of the concepts listed above,
related in definite ways. It is, for instance, obvious that in the concept of a
material particle moving through space along a continuous path, persisting
through time and obeying certain laws of dynamics, reference is being made
to the other concepts just mentioned; thus the changes of these other con-
cepts do inevitably affect the concept of particle and vice versa. It is impor-
tant to keep in mind that the crisis for the corpuscular-kinetic models means
not only a crisis for the concept of particle; it involves a set of all the corre-
8 MILIC CAPEK

lated changes in all the other constituent concepts. Thus it is impossible to


discuss the transformation of the concept of matter and, one might even say,
the elimination of the concept of particle without referring to all the other
correlated conceptual changes. Only such an approach can help us to place a
new concept of matter in a proper perspective and to grasp the full meaning
of the twentieth-century revolution in physics. Nothing in my opinion is
more dangerous to a full understanding of this topic, than a discussion
of the change in each individual concept, in isolation from the changes in
the whole conceptual framework. Yet, I ·am afraid this is what frequently
happens, as I shall later illustrate, using some concrete examples.
As I said before, the electron theory of matter in its first phase seemed to
represent the successful culmination of centuries of effort to interpret physi-
cal reality in a corpuscular-kinetic fashion; yet, it is interesting that the first
signs of the inadequacy of the classical concept of corpuscle began to appear
at the same time. The atom was no longer regarded as indivisible, but as com-
plex; but this in itself was hardly a threat to the concept of corpuscle, since
its constituent intra-atomic parts, whether those composing the nucleus, or
orbiting around it, were still regarded as particles; they were merely the old
solid atoms of Dalton, but on a much smaller scale. They were regarded as
tiny spheres, possessing a certain mass and consequently being indestructible
- a simple consequence of the law of conservation of matter; they seemed to
be the true ultimate units of matter. But there were some complications
which, from the rigorously mechanistic point of view, appeared as serious
flaws. First, they were viewed as the basic units of electricity: the electrons
with a negative and the protons with an equal positive charge. Now there
were two questions involved here, about the relationship between the mass of
the particles and their charge, and about the dual character of these charges.
Neither of these questions could be answered satisfactorily. Were the mass
and the charge two irreducible attributes of the particle? This idea hardly
appealed to mechanistic taste, which was always idiosyncratically reluctant to
concede any irreducible diversity or even duality in nature. Or was mass
merely a manifestation of electric charge? This latter view was fashionable
among some physicists at the turn of the century, and as late as 1914, Jean
Perrin suggested in his now classical book Les atomes that perhaps the whole
inertial mass of the electron may be of electromagnetic origin. 9 This view was
suggested by the increase in the inertial mass of the electron which follows
from Einstein's relativistic mechanics, but was originally interpreted in the
terms of classical electromagnetic theory as a result of the reaction of the
electromagnetic field to the motion of the electron. From this point of view,
PARTICLES OR EVENTS? 9

the electrons themselves and the elementary particles in general "are nothing
but condensations of the electromagnetic field" (Einstein).lo
But would this mean the reduction of matter to electromagnetism?
Einstein continues the passage quoted above as follows: " ... our conception
of the cosmos recognizes two realities which are conceptually quite indepen-
dent of each other even though they may be causally connected, namely
gravitational aether and the electromagnetic field ... ". This would mean that
not all manifestations of matter are reducible to electromagnetism. It is true
that the reference to aether would seem to give another chance to the
mechanist provided he could construct a successful mechanistic model of the
medium which would account for the transmission of electromagnetic waves
and also, it is to be hoped, of gravitational interactions, and in which the
elementary 'particles would merely be local structural complications. Such
was the hope of all mechanistic theories of aether from that of Huygens up
to William Thomson. This would in fact amount to an operation in just the
opposite direction, namely the reduction of the whole of physical reality to
the mechanics of the aetherial medium. If the structure of this medium were
grain-like, as Huygens originally suggested and as a number of physicists still
believed at the tum of the century, the concept of ultimate corpuscular units
might still be saved, but their dimensions would be incomparably smaller than
those of the electron; thus Osborne Reynolds (1903) estimated a corpuscular
radius equal to the order of the 1O-18 cmY
But such hopes were already obsolete at that time, and even more so by
the time Einstein gave his lecture about aether. Even prior to the special
theory of relativity, the inadequacy of the corpuscular-kinetic models mani-
fested itself in the repeatedly frustrated efforts to construct a satisfactory
mechanical model of aether. The final blow to such hopes was the negative
result of Michelson's experiment which divested aether of even the most
elementary kinematic properties; it had to be neither at rest nor in motion,
otherwise there would be inconsistence with the fact of the constant velocity
of electromagnetic waves. For all practical purposes the idea of aether wad
dead and if Einstein was still willing to retain the original word, its meaning
was so thoroughly different from the original one that an entirely different
term should have been invented for this purpose. Today only a few people
still speak of aether and hardly any of 'aether particles'.
But the term 'material particle' is still very much alive, even though, as I
am going to argue, it is no less inappropriate than the term 'particle of aether'.
Let us return to the original planetary model of the atom which at first
appeared to be another triumph of the corpuscular-kinetic view of physical
10 MILIC CAPEK

reality. Without the help of aether the relationship between the mass of the
particles and their charge as well as the duality' of charges remained myste-
rious. Fifty years ago it was already clear to Hermann Weyl12 that to imagine
the charge as 'sticking' to the rigid electrons would be nothing but a grotes-
que naivete. But this was not the only difficulty. Another was an apparent
lack of proportionality between mass and volume - a proportionality which
was one of the cornerstones of classical atomism. Thus although the proton is
nearly two thousand times heavier than the electron, its radius is of the same
order of magnitude, i.e., lO- 13 cm; and this is apparently true of other
elementary particles. The third difficulty appeared when Niels Bohr intro-
duced his quantification of electronic orbits in 1913; it then became abun-
dantly clear that the alleged analogy between a macroscopic planetary system
and the atom was altogether deceptive. The so-called 'forbidden zones'
between discrete electronic orbits clearly did not have any macroscopic coun-
terpart. Even more seriously, they seemed to contradict both the homo-
geneity of space and the continuity of trajectories on the microphysical scale.
Russell's 'axiom of free mobility' which characterizes not only Euclidean
space, but also all homogeneous spaces (i.e., those with constant curvature)
ceased to be applicable to microphysical space. Related to this was the diffi-
culty of applying a classical spatio-temporal analysis to the so-called
'quantum jumps'. If the electron were really a corpuscle, then its passage
from one orbit to another should be along a continuous path, no matter how
short, from a point in one orbit to a point in another orbit. But from the
beginning it was if not obvious, then at least very probable that such a transi-
tion from one orbit to another should be regarded as an indivisible jump
within which neither spatial nor temporal subintervals could be discerned.
Thus even in the early phase of the development of quantum theory, doubts
began to emerge about the classical continuity of space, time and motion;
Poincare before 1912, followed by Whitehead in 1920 considered the possi-
bility of an 'atom of time' or a 'quantum of time'; this was soon followed by
a host of speCUlations about the discrete nature of space, time and motion. 13
But without spatio-temporal continuity of its path, the identity of the particle,
i.e., its identifiability at different points, in space and at successive instants
of time is impossible to maintain and the concept of corpuscle itself loses it
meaning.
There was another large group of facts, discovered and interpreted
by relativity theory which suggested an equally profound revision of the
concept of particle. The special theory fused together two concepts tradi-
tionally distinct - mass and energy. In classical physics, mass and energy had
PARTICLES OR EVENTS? 11

always remained distinct and one could exist without the other: thus a particle
at rest was devoid of energy while the radiation energy was regarded as mass-
less. Even when a particle was in motion, its mass remained unaffected by the
kinetic energy associated with it. It is different in relativistic mechanics: an
increase in kinetic energy involves an increase in mass - in other words, the
mass of a particle in motion is greater than that when it is at rest; this fact
was amply confirmed on the microphysical scale, for bodies whose velocities
were significantly close to the velocity of light. Similarly, allegedly 'disem-
bodied' electromagnetic radiation has a certain inertial mass, though small,
and thus exerts a certain pressure - a fact verified even prior to the advent of
the special theory, by Lebedev in 1900. The concept of mass was thus genera-
lized, but at the same time was clearly divested of its original intuitive (Le.,
corpuscular) connotation.
Other consequences of the special theory or, more specifically, of the
relativistic 'fusion' of mass and energy, point in the same direction. Einstein's
equation E = mc 2 means that every increase or decrease in energy involves an
increase or decrease in the corresponding inertial mass. Thus the total mass of
a material aggregate is no longer equal to the sum total of the masses of the
particles of which it is composed, as was true in classical physics, and as is still
approximately true in our daily macroscopic experience. It is either decreased
or increased as energy is either absorbed or released in the process of aggrega-
tion. Thus, strictly speaking, there is a loss of mass when one mole of CO 2 is
formed, since 94.052 calories of energy are liberated; while the absorption
of 21.600 calories in the formation of NO results in an increase in mass. It is
easy to see that in such reactions, as in every macroscopic chemical reaction,
the calculated mass effects, whether positive or negative, are too minute to be
experimentally detected. This is why they escaped detection even in the
accurate and repeated experiments of Hans Landolt at the turn of this
century, the results of which were hailed as a definitive confirmation of the
law of conservation of matter (or more accurately, of mass). But what
Landolt proved was only that there are no relative variations in weight to the
order of 10- 6 , while the relativistic variations in mass in ordinary chemical
reactions amount to less than 10- 13 of the total mass involved.
But the situation is different when one considers aggregate formations on a
nuclear scale of magnitude. One of the most well-known instances is the mass
effect resulting from the formation of the nucleus of a helium atom, consisting
of two protons and two neutrons; while the mass sum of all the components
is 4.03302 atomic mass units (a.m.u.), the mass of the compound nucleus is
less - only 4.00280 a.m.u .. In other words, approximately 0.03 atomic mass
12 MILIC CAPEK

units have 'disappeared' or, more exactly, have been converted into the
binding energy of the nucleus. Similar mass decrements have been found for
other elements; these increase with atomic number up to 0.238 a.m.u. for
uranium. The frightful technological application of the energy released by the
so-called 'annihilation' of mass is generally known. The opposite process, i.e.,
the 'materialization' of energy takes place in so-called endergic reactions where
energy is absorbed instead of being released; it is, especially striking in the
reactions between elementary particles of very high energy when some of
their kinetic energy is converted into the rest mass of a new particle. The
creation of a 1T-meson by the interaction of two high energy protons is an
example of this kind. The first instance of this process was observed in
1932 with the discovery of the positive electron by Anderson; in this case it
was the energy of high frequency radiation interacting with a heavy nucleus
which was for the most part 'materialized' into the rest mass of two oppo-
sitely charged electrons while the excess of the original energy survived in the
form of their kinetic energy. The opposite process of 'dematerialization' was
discovered at about the same time; in fact, the reason why the positive elec-
tron was discovered so late was its extremely short life; its dematerialization
occurs after 10-8 sec when it encounters a normal, i.e., negative electron,
and they both disappear in a puff of high frequency radiation.
The observed variability of the mass of the elementary particles as well as
their creation and annihilation are the most serious threats to the applica-
bility of the concepts of corpuscle on the microphysical scale. For is it
meaningful to apply the term 'corpuscle' - the most salient traditional fea-
tures of which were immutability and everlastingness - to mesons, some of
which 'last' only 10- 16 sec? Is not the term 'event' a more appropriate name
for such evanescent entities? This question becomes even more pressing when
we realize that materialization and dematerialization of particles is not an
exception, but rather a rule. Even such a 'solid' particle as the neutron decays
in twelve minutes, while the more massive hyperons disintegrate in one
hundred millionth of a second. At first glance these processes look very
different from those of creation and annihilation of particles; the 'disintegra-
tion' and 'decay' are perfectly meaningful within the corpuscular-kinetic
scheme - all we have to assume is that the so-called unstable particles are
really composite and not indivisible, and that radioactive decay is nothing but
a drifting apart of the constituent particles which were originally closely
packed together. But as Niels Bohr observed as early as 1939,14 to assume
that beta-particles pre-exist in the nucleus, from which they are then ejected,
is as naive as to believe that photons pre-exist in the atom prior to their emis-
PARTICLES OR EVENTS? 13

sion. On the contrary, according to Bohr, instead of expulsion of the electrons


from the nucleus, we should speak of their creation; they are created in a
sense similar to that in which photons are created during their emission. The
classical idea of the nucleus being made up of smaller, juxtaposed and closely
packed subparticles, although appealing to our pictorial imagination, simply
fails to represent the true microphysical situation.
This can be shown quite convincingly by analyzing the radioactive decay
of the neutron. If a neutron were really a close combination of a proton and a
negative electron, and if its disintegration could be regarded as a separation of
these two constituent parts, then the opposite process - the transformation
of a proton into a neutron and a positive electron would be impossible. Yet
this process does take place. Now this remains completely unintelligible
within the classical corpuscular-kinetic scheme: either the neutron is an
aggregate which disintegrates into its constituents - the proton and the
negative electron - or the proton is complex and the neutron and the positive
electron are its parts. But we cannot have it both ways since this would not
make any sense: a particle which would be a fragment of another particle
cannot contain the same particle as its own part. Here we are witnessing a
complete failure of a strictly corpuscular or configurational model of the
nucleus. Neither proton nor neutron are configurations in the classical sense
of being composed of the more basic, closely packed, juxtaposed units,
actually existing prior to their separation. As Otto von Frisch observed, the
situation is even worse when we consider so-called 'strange particles'. "The
,,-meson can change into two pions or into three pions or into a pion, a muon,
and a neutrino, or in several other ways. If we assume that ,,-meson is com-
posed of, say, three pions, then we cannot understand the other modes of
break-up." He then logically concludes that "the very idea of compositeness
must be left behind if we want to understand the subatomic particles."
Heisenberg arrived at the same conclusion. 15
There are, of course, ways to make these paradoxical results appear
intellectually more palatable. For instance, it is true that the materialization
of the particles is not creatio ex nihilo, creation out of nothing, nor is their
'dematerialization' their absolute annihilation; the conservation laws of mass-
energy are not formally violated and one may be allowed to say that the mass
of a couple of the created electrons 'virtually' pre-existed in the mass-energy
of the radiation from which they originated. Thus Hans Reichenbach
exaggerated when he called these phenomena "causal anomalies" .16 But, as I
mentioned above, this formal preservation of the constancy of mass is
obtained by the generalization of the concept of mass itself, that is, by
14 MILIC CAPEK

merging it with the concept of energy - but only at the expense of divesting
it entirely of its original, i.e., corpuscular connotation. For it is obvious that
such paradoxical fusion of mass and energy implies consequences which
appear grotesque and absurd as long as we retain the corpuscular-kinetic
framework. In such a framework there is no place for genuine 'virtualities' or
potentialities' since its conceptual components are precisely the bits of homo-
geneous matter assumed to exist actually, i.e., actually occupying various
positions in actually existing space. To speak of 'virtual particles' as present-
day meson physics does is hardly anything more than a concession to old
intellectual and linguistic habits. As Emile Meyerson observed, the notion of a
'potential' or 'virtual' state is a linguistic device to preserve the identity and
uninterrupted continuity of the object in time, by assuming that it somehow
continues to exist even if it apparently disappears. I7 Thus its 'potential' or
'virtual' existence during the intervals of its unobservability guarantees its
persistence of identity in time. Only in this way can the human intellect (in
its classical form, we must add today) eliminate the emergence of genuine
novelties and reduce all changes to a mere reshuffling of permanently existing
particles. Thus in most instances, especially in the classical era, the term
'virtual' is merely a cover-name for 'actual', and a potential entity is under-
stood as a hidden actuality. Hence the persistent hopes to interpret potential
energy as the kinetic energy of actually moving, invisible particles - hopes
which can be traced from Christiaan Huygens to Herbert Spencer, that is, to
the very end of the 'classical era'.18 Fortunately, hardly any physicist today
maintains the 'virtual' pre-existence of the mass of the pair of electrons in the
sense of their actual, corpuscular pre-existence; in fact, the adjective 'virtual'
which physiCists join to the noun 'particle' is intended as an explicit warning
not to take the word 'particle' in a literal sense. But in such a situation, would
it not be better to drop the word 'particle' or 'corpuscle' altogether, precisely
because it is so loaded with misleading associations which no qualifying
adjective can successfully eliminate?

III. POPPER-LANOE'S DEFENSE OF THE REALITY OF PARTICLES


I could go on accumulating other examples of the inapplicability of the
strictly corpuscular concept, but new examples would not be basically
different in their import from those already mentioned. Furthermore, at this
stage of my exposition I can expect the following, rather impatient remark
from either a physicist or a philosopher of science: "All you are trying to
prove is that the term 'corpuscle' or 'particle' cannot now be taken as having
its original, literal meaning. But everybody knows this since everybody is
fully aware of the fact that corpuscular character is only one aspect of
PARTICLES OR EVENTS? 15

physical reality and that there is another aspect, equally essential and comple-
mentary to it, corresponding to the wave nature of matter. The inadequacy of
the classical concept of particle is due to the fact that the wave aspect must
also be taken into account - and that everybody knows."
Well, not everybody. There are serious and outstanding thinkers who still
claim that the wave aspect is merely secondary and, so to speak, apparent and
thus unrelated to the fundamental physiCal reality which consists of particles.
This is, for instance, the view of Karl Popper and Alfred Lande who claim
that Heisenberg's uncertainty relations have been habitually misinterpreted
since an attentive analysis of them will disclose that they do not impose any
definite limit on the precision of simultaneous determination of the position
and ihe momentum of a particle and thus are compatible with the reality of
particles. 19 This view is diametrically opposed to the view of such widely
different thinkers as Arthur Eddington, Philipp Frank and Max von Laue who
argue that there is an objective limit to the precision of our measurements
and that the impossibility of measuring exact position and exact momentum
in conjunction can mean only one thing: that such a conjunction simply does
not exist in nature. 20 Since such a simultaneous conjunction of position and
momentum is nothing but a particle itself, the denial of the reality of such a
conjunction is equivalent to the denial of the reality of particles.
This would follow as a direct consequence of an objectivistic interpretation
of Heisenberg's principle, the correct name of which, in such a case, should
be Indeterminacy rather than Uncertainty, principle. If this view is correct-
as I have tried to show elsewhere 21 - then the principle of indeterminacy
would be the final coup de grace to the concept of particle or rather to its
applicability, the many inadequacies of which I have tried to point out.
Now it would be unfair to discuss Popper and Lande's argument in the
limited space which remains available for me; this really requires a separate
paper. But allow me to make some modest and, I hope, relevant remarks.
Popper is justifiably proud to be, as far as his view is concerned, in good
company with such men as Einstein, Louis de Broglie, Max Born, Lande and
Bohm; but it is only too clear that this company is rather heterogeneous. Thus
Einstein and de Broglie's rejection of the usual interpretation of Heisenberg's
principle was largely inspired by their commitment to determinism; to some
extent, though not entirely, this is also true of David Bohm, but certainly not
of Lande to whose views Popper's are closest. They both assert the physical
reality of particles together with indeterminism. This sounds strange only
if we forget that according to both Popper and Lande, not only quantum
physics, but classical physics as well must be regarded as indeterministic or at
least not rigorously deterministic - a rather paradoxical view and histOrically
16 MILle CAPEK

incorrect at that. But this is less important in the context of our discussion
than the fact that both Popper and Lande are apparently not ready to face all
the consequences of their affirmation of the reality of particles.
If particles are truly physically real, i.e., if they have at each instant an
exact position and an exact momentum, then they should move along con-
tinuous trajectories both outside and inside the atom, and even inside the
nucleus. This would mean a return to the old, naive planetary model of the
atom; instead of energy levels, we would have electrons moving continuously
in circular or elliptical orbits around the nucleus; it would mean that an
electron would literally jump from one orbit to another and during this jump
would move continuously - no matter how quickly - through the inter-
vening zone between the orbits. Finally, it would mean - if Popper and
Lande really mean what they say - than even within the nucleus the particles
would move continuously, possessing at every instant an exact position and
an exact momentum, no matter how quickly their positions and velocities
might change. Few, if any of the defenders of corpuscular models go so
far, even though it would only be consistent for them to do so. Take, for
instance, the spin of the particles. The only meaningful way to integrate it
into a corpuscular framework is to interpret it as a rotation of tiny spheres.
But as has been shown long ago, this would imply rotating velocities ex-
ceeding the velocity of light, i.e., incompatible with relativistic mechanics?2
On the other hand, to give up this naive interpretation borrowed from the
mechanics of macroscopic bodies, while retaining the corpuscular models, is
hardly satisfactory; it results in a schizophrenic mixture of incompatible
epistemological attitudes, half way between Kelvin and Dirac: an abstract,
non-intuitive property is incongruously grafted on to the pictorial image of a
tiny sphere. Those are not the only difficulties. A far more serious - in fact
insurmountable - difficulty is to interpret the undulatory character of matter
in terms of consistent corpuscular models; but more about this later.
Why then are the majority of rank-and-file physicists together with some
philosophers of science so strongly committed to the reality of particles?
Because they are sincerely convinced that the empirical evidence for their
existence is overwhelming and irrefutable. They point out that the discrete
tracks in a Wilson Ooud Chamber as well as discrete impacts on the spinthari-
scope screen or on a photographic plate cannot be interpreted in any other
way. Lande himself regards these facts as 'directly supporting' the objective
reality of particles. Yet it has been pointed out long ago that this is not
necessarily SO.23 What looks like a 'continuous corpuscular trajectory' in the
Wilson Chamber under a magnifying glass appears as a discontinuous series of
PARTICLES OR EVENTS? 17

water drops separated by irregular intervals, arranged along almost straight


lines in the case of fast-moving and heavy particles, while the 'tracks' of
slow-moving electrons are sinuous or broken lines_ Each droplet has condensed
around an ion and thus what appears to us as the 'continuous track of a parti-
cle' is nothing but a discontinuous series of water drops reflecting the light.
But this is not alL Each fast-moving electron successively strikes different
molecules of gas; it ionizes each of them by tearing away a peripheral electron
from it, so that the rest of the molecule acquires a positive charge. (For
the sake of argument, I am deliberately using. the original corpuscular lan-
guage of Bohr's model.) Are we now sure that it is the same electron which
continues along this track? Or is it an electron. which was driven from the
ionized molecule? Obviously, no answer is possible. In other words, the
allegedly 'direct' empirical evidence for the existence of particles guarantees
neither the identity of the particle at different points along its alleged trajec-
tory nor the continuity of its path - the two features which are logically
inseparable and without which no consistent corpuscular model is possible_
TIlis clearly shows how in thinking about the 'microcosmos' we are easily
misguided by false macroscopic analogies. When we see periodical flashes of a
firefly in the darkness, we assume that the same firefly continues to exist
between its flashes, and we can easily verify this; but, as Professor Mar-
genau24 asked, are we justified in using such a 'firefly model' in microphysics?
Unlike a macroscopic object, an electron's motion cannot be followed with-
out being lost from our sight; Mill's defmition of ,matter as the 'permanent
possibility of sensations', applicable to a firefly and the macroscopic objects
generally, fails when we try to apply it to micro-objects. But if the trajectories
of the particles are not continuous in the sense that the particles have posi-
tions and velocities (if we disregard for a moment Heisenberg's uncertainty
relations) only at certain instants and not between, we give up their claim to
existence in everything but name. It would be an abuse oflanguage to apply
the name 'particle'to a string of spatially and temporally disconnected events.
We have just mentioned Heisenberg's uncertainty principle; it is obvious
that to deny it or at least to deny its objectivistic interpretation is essential
for those who, like Popper, retain the reality of particles. For if we accept it,
we concede that not only can no particle be observed twice (as Schrodinger
correctly observed 25 and as follows from the analysis above), but that it
cannot even be observed once since there is no such thing as a conjunc-
tion of exact velocity and exact position, in nature. It is true that if we
regard the principle as a mere limit to our technique of observation (and,
according to Popper, a limit which is not unsurpassable), then there may still
18 MILIC CAPEK

be hope for the existence of hidden corpuscular structure at the sub-quantum


level. Whether such hopes are borne out is another matter; the circumstantial
evidence against the divisibility of the quantum of action is so far overwhel-
ming. But it is interesting that not all those who share such hopes accept a
corpuscularian philosophy of the Popper-Lande kind. Thus David Bohm
doubts that the inner structure at the sub-quantum level "is made of still
smaller particles":

For the field theory describes all motion in the quantum-mechanical domain in terms
of 'creation' and 'destruction' of elementary particles. Thus, if an electron is scattered
from one direction of motion into another, this is described as the 'destruction' of
the original electron, and the 'creation' of another electron moving in the new direction.
Hence, there is, in this theory, no particle which permanently retains a fixed identity
as a particle. Indeed, if one looks more deeply into the field representation of the
motion even for a free particle one discovers that its motion is described mathematically
as a destruction of the particle at a given point and its creation at a closely neighbouring
point. Thus the motion is analysed as a series of creations and destructions, whose net
effect is to continually displace the particle in space. 26

Bohm concludes that if the notion of a permanent particle fails on the quan-
tum level, a fortiori it is inapplicable on the hypothetical sub-quantum level:
if there are such micro-microparticles, they are "always forming and dis-
solving" sO,that their precise positions and momenta would have little signi-
ficance. I think that still to call such quickly dissolving structures 'particles'is
an excessive concession to traditional linguistic habits.
In a similar way, J.-P. Vigier defines his hypothetical particle as "an
average excitation of a chaotic sub quantum-mechanical level of matter;
similar in a sense to a sound wave in the chaos of molecular agitation". 2 7
Philosophically, Bohm and Vigier's views belong to the same category as
those of Schrodinger and Einstein: the 'particles' are merely temporary
perturbations in the continuity of the spatio-temporal field - a tradition
whose roots go back to William Thomson, aether theories and eventually to
Cartesian physics. Although classical atomism was radically opposed to
Cartesianism and its posterity as far as the status of particles was concerned,
it shared with Cartesianism two fundamental theses: it accepted the con-
tinuity of space and time (only matter is discrete, not space, nor time) and,
most importantly, rigorous determinism.
Only on this last point is there a significant and rather strange difference
between classical atomism and its new Popper-Lande version. As already
mentioned, their commitment to the reality of particles does not prevent
PARTICLES OR EVENTS? 19

them from being indeterminists - if their indeterminism is genuine. Only


once in Western intellectual history has such a combination of corpuscularism
and indeterminism appeared. Its only (still famous) instance was the post-
Aristotelian atomism of Epicurus and Lucretius; but it was generally re-
cognized that in their system the idea of indeterminacy was incongruously
grafted onto a mechanistic scheme. From Cicero to Bergson, philosophers
have dismissed and ridiculed it; when Gassendi tried to rehabilitate Epicurus's
atomism, it was atomism without the spontaneous swerve of the atoms. When
some late nineteenth-century thinkers favored the objective status ofindeter-
minacy in the physical world, they tried to dissociate it from its original cor-
puscular-kinetic framework: such was the case with Renouvier's neomona-
dism, Boutroux's contingentism and Peirce, Berg'!on and Whitehead's panpsy-
chism. Very few historians found kind words for Lucretius's indeterministic
mechanism: it is true that Emile Meyerson, always looking for historical ante-
cedents, regarded the Lucretian clinamen as a remote ancestor of quantum
indeterminacy and that Dirac probably unintentionally paid a compliment to
the author of De rerum natura when he spoke, either jokingly or seriously, of
liberum arbitrium naturae. 28 But there is hardly any question that when an
attempt is made to combine mechanism and indeterminism, something similar
to Epicurean atomism must result. For the only way to smuggle indeter-
minacy into a corpuscular-kinetic scheme is to endow the particles with the
capacity to choose their future positions and future velocities from a certain
range of possible positions and possible velocities. Popper's theory of particles
endowed with 'propensities', when placed in the wider perspective of the
history of ideas, appears as a new, highly sophisticated version of the Lucre-
tian doctrine of clinamen. It is hardly surprising that it faces the same diffi-
culty: how to integrate a non-mechanistic element of 'propensity' or 'poten-
tiality' (Popper uses both terms)29 into a corpuscular, i.e., mechanistic
scheme.
Besides this major philosophical difficulty there is another which only took
on a more concrete form as a result of twentieth-century physics: how do
we integrate the undulatory character of matter with the corpuscular model?
According to both Popper and Lande, the primary reality, which consists of
particles and their 'wave-like character', is nothing but the way they are distri-
buted in space. There is thus no duality of particles and waves; this duality
was already removed by Max Born, and Lande only regrets Born's occasional
reversion to dualist language,30 language which also prevails in the majority
of the textbooks. But then why are the particles distributed in a wave-like
fashion? Where does the spatial periodicity of their interference ring'! come
20 MILle CAPEK

from? Both Popper and Lande reject the attempts to explain this periodicity
in terms of some unitary wave theory, whether one speaks of Schrodinger's
original theory of particles as high crests of real, substantial waves, or of the
later theories of de Broglie, Bohm and Vigier. All such theories are deter-
ministic as well as anticorpuscularian, in the sense that particles are regarded
as mere temporary products of the continuous field whose vibratory distur-
bances are fully describable in a classical deterministic fashion. Such views are
obviously diametrically opposed to Popper-Lande's insistence on the substan-
tial reality of particles as well on the irreducible character of the statistical
laws. But this leaves the basic question mentioned above unanswered: why do
electronic impacts form interference rings, instead of being distributed in a
classical, random, non-periodical fashion?
One can feel considerable sympathy for the Popper-Lande criticism of the
orthodox interpretation of quantum theory, with its positivistic, phenomena-
listic and even subjectivistic overtones. Popper rejects Jeans's view that the
probability waves are 'waves of knowledge', and rightly criticizes the oscilla-
tion between subjectivistic and objectivistic interpretation of probability
which is so characteristic of the Copenhagen school. His real merit is that he
stresses that physical indeterminism, i.e., the primary and irreducible character
of the statistical laws, does not necessarily imply epistemological idealism; he
accepts Einstein's epistemological realism without accepting his determinism.
This is why he insists that the waves of propensities are "physically real",
being "objective relational properties of the physical world." (Lande is more
radical than Popper since he insists on the exclusive reality of particles and
deprives probability of any physical, 'kickable' statUS. 31 But since he also
rejects de Broglie and Bohm's theory of pilot waves, the physical status of his
propensity waves remains obscure. The obscurity is only increased by his
claim that quantum theory is no more mysterious than any other game of
chance and that "all apparent mysteries would also involve thrown dice, or
tossed penny exactly as they do electrons".32 (His italics.) Such comparison
of the behavior of the electron with that of a macroscopic object is epistemo-
logically untenable, but for both Popper and Lande it follows from their
equally untenable denial of any basic distinction between classical and con-
temporary physics. Their tacit assumption is analogous to that of Einstein: in
the same way that Einstein dogmatically assumed that epistemolOgical realism
is possible only on a strictly deterministic basis, Popper and Lande believe
that it is possible only on a corpuscular basis. But does the attitude of
epistemolOgical idealism necessarily correspond to a return - no matter how
disguised and sophisticated - to Epicurus and Lucretius?
PARTICLES OR EVENTS? 21

IV. A CASE FOR EVENTS

There is another alternative open to us besides those mentioned above: that


the basic physical entities are neither particles nor waves in their usual macro-
scopic sense, but imageless events whose behavior, when it is communicated
to us by very complex artificial arrangements, reminds us alternately of the
behavior of macroscopic particles or periodical disturbances in elastic media.
Enough has been said about the inadequacy of the concept of particle; the
intuitive image of wave as a periodical disturbance in the three-dimensional
aetherial medium proved to be inadequate for the same epistemological
reasons. Thus 'we can afford to be fairly brief in our survey of the main
difficulties with the concept of aetherial medium.
The first signs of its inadequacy were the enormous difficulties encoun-
tered in trying to construct a consistent and empirically satisfactory model.
These difficulties had become insurmountable after Michelson's experiment
which deprived aether of its last intuitive properties: aether was deprived even
of the most fundamental kinematic properties, for it could not be either at
rest or in motion. Such difficulties increased enormously with Planck's dis-
covery of the quantum character of radiation which showed that the propaga-
tion of light cannot be exhaustively described· in terms of continuously
spreading waves. Thus a highly unsatisfactory 'wave-particle dualism' was
introduced into physics - unsatisfactory as long as 'particles' and 'waves'
were understood in their original, intuitive, macroscopic sense. It is generally
known that quantum theory was in no sense a simple return to the old
corpuscular theory of light ; nor could the 'waves of matter' (les ondes materiel-
les, Materiewellen) discovered by de Broglie, be interpreted in a realistic sense
as periodical disturbances in a hypothetical sub-aether, as Schrodinger
originally believed. Intuitive models of waves are thus as unsatisfactory as
intuitive models of particles and it is rather interesting that de Broglie's dis-
covery of the vibratory nature of matter represented another 'coup de grace'
for the traditional pictorial image of corpuscle. For no effort of imagination
can integrate into one single self-consistent model the image of particle and
that of vibration.
The last statement hardly needs any extensive comment: a particle can
vibrate, but it cannot consist of vibrations. If we say that it does consist of
vibrations, being, as was originally believed, that region of the 'sub-aetherial
field' where the amplitudes of the vibrations are higher than in its neighbour-
hood, then it ceases to be a particle in the original sense because it loses its
substantiality. Conceptually, within the classical intuitive framework, 'parti-
22 MILIC CAPEK

cle' and 'wave' remain irreducibly different and can be associated only in an
external way. This is why physicists speak of the de Broglie wave length being
'associated with' a particle according to the formula X = h/mv. Strangely
enough, an equally fundamental equation mc 2 = hf(m = mass, c = velocity of
light, h = Planck's constant, f = frequency) is hardly ever commented upon.
What is the meaning of this intrinsic vibrational frequency f, once we give up
the intuitive and naive tendency to interpret it as a locomotory oscillation of
some imaginary 'sub-aetherial' corpuscles?
On this point, the view of the particle as a string of imageless events seems
to me the only one that is free of the epistemological crudities of visual
mechanistic models, in particular of the intolerable conflict between the two
classical images. It presupposes a radical revision of classical habits of thought
and as such it cannot be welcomed by those who insist on the 'Cartesian
clarity' of physical models. In particular, it presupposes a negation of the
infinite divisibility of time as envisioned by Whitehead and, before him, by
Poincare and Bergson. It takes into account the two most philosophically
significant innovations of relativity - the conjunction of space with time and
the elimination of the distinction between time-space and its physical content.
Thus to say that particles are 'successions of spatio-temporal pulsations' is not
a mere figure of speech; as Whitehead pOinted out shortly after de Broglie's
discovery, "when we translate this notion into the abstractions of physics, it
at once becomes the technical notion of 'vibration,.,,33 The Significance off
in the equation above is clear: it designates a number of elementary events
constituting the duration ofa particle per unit time. These events are precisely
the spatio-temporal minima whose extents vary according to the nature of par-
ticles, although, in the light of present evidence, it is never significantly below
the limits postulated by the chronon and hodon theories.
A physicist will probably object that the view proposed above is too general
and qualitative to be satisfactory. Indeed, it is; it is not a model in the usual
sense. You may even call it a philosophical guess. But the task of a philosopher
of nature is not to prescribe directives too specific to the development of
science; such attempts have always ended catastrophically, as the fate of the
German Naturphilosophien, in particular those of Schelling and Hegel, shows.
His task is to discern the direction in which the sciences are moving; in this
particular case, to take into a synthetic account the cumulative evidence of
the inadequacies of the corpuscular, and more generally, mechanistic models.
By focussing our attention on conceptual blind alleys, a philosopher of nature
can open the way for new channels of thought which appear implaUSible only
because of our adherence to traditional modes of thinking. Thus the implausi-
PARTICLES OR EVENTS? 23

bili ty of the view that matter consists of vibrations or -less intuitively - of a


succession of exceedingly short events is due to the fact that in all classical
models, the particle, besides filling a certain tiny volume of space, also fills
continuously and without interruption any interval ofits duration, no matter
how small. In Whitehead's words, ''if material has existed dUring any period,
it has equally been in existence during any portion of that period. In other
words, dividing the time does not divide the rnaterial".34 In a more ordinary
language, the particles persist uninterruptedly through time. (I do not see any
evidence that either Popper or Lande depart from this classical view.) In the
space-time diagram such particles are represented by narrow four-dimensional
tubes of infinite length because of their permanence and indestructibility. It
is this model which exerts an inhibitory influence on our minds and which
makes the vibratory model of matter so difficult to accept.
But the situation changes when, again in Whitehead's words, ''we consent
to apply to the apparently steady undifferentiated endurance of matter the
same principles as those now accepted for sound and light". He is careful to
add that such a vibration ''is not the vibratory locomotion: it is vibration of
organic deformation". It is not a vibration of something like the vibration of
the particles of air in sound or that of some mythical aethereal or sub-aethereal
particle. For this reason, instead of using Whitehead's highly metaphorical
neologism, I prefer to speak of a 'succession of events' or a "succession of
spatio-temporal deformations' 'or 'pulsations'; I am aware that these are also
neologisms, but they are less tainted by misleading associations. (To every per-
son acquainted with physics, the word 'vibration' inevitably suggests a spatial
periodical displacement, of something around a position of equilibrium - and
thus steers his imagination back to the discredited notion of some substantial
particles.) But in this context we immediately see one concrete advantage of
our model which at first appears so unsatisfactorily general. As long as the
particles are viewed as substantial entities, persisting continuously through
time, their interference and diffraction patterns remain unintelligible; highly
artificial assumptions must be made to explain why they are scattered only in
certain specific directions to simulate the wave-like behavior. On the other
hand, if the electrons and micro-particles in general are intermittent 'entities'
or events, their interference becomes in principle possible. Two particles, be-
cause of their permanence, can never merge together nor cancel one another;
two events can. Such cancellation clearly occurs in the minima of interference
rings, and, more spectacularly, as Otto von Frisch pointed out, in a 'corpuscular
version' of Michelson's experiment, where two electron beams cancel each
other.3s This does not mean that the electron consists of waves, but only that
24 MILIC CAPEK

it has intrinsic vibratory, i.e., periodical structure. Now the essence of any
interference is the presence of the periodicity in time which produces the
periodicity in space, i.e., the alternating regions of maxima and minima.
But the most decisive argument for the view stated above is an epistemolo-
gical one: namely the argument based on the genetic or biological theory of
knowledge. I concede that such an argument can interest the rank-and-file
physicists only indirectly, but its cogency is recognized by some philosophi-
cally-minded physicists and some philosophers of science. Only the barest out-
line can be given here; besides this, I refer the reader to my previous articles
on Mach, Bergson and Reichenbach and to the whole first part of my book
on Bergson. Briefly stated, this theory is as follows: the cognitive functions of
the human mind are not static entities, but they are results of evolution; in
their present form, they are the result of a long adaptive process by which the
human mind adjusted itself to external reality. According to the older biologi-
cal theory of knowledge, proposed by Herbert Spencer and upheld by
Helmholtz, Mach and even Poincare, this evolutionary process was completed,
at least in its essential features; human intelligence in its classical form was
regarded as the final and culminating phase of this adaptive process in the
sense that the traditional two-valued logic, Euclidean geometry and finally
classical mechanics were believed to represent a true, adequate and in their
general features complete representation of reality. (I suspect that both
Popper and Quine are fairly close to this view.) It is interesting that this older
version of biological epistemology agreed in one important aspect with Kant.
For Kant, any future departure from the Newtonian-Euclidean picture of
reality was logically excluded by the rigid and unchanging a priori structure
of the human cognitive apparatus; for evolutionary positivists of the last
century it was excluded because the Newtonian-Euclidean character of our
intelligence is the final adaptation to objective reality; any departure from it
would be a step backwards, an epistemological regression.
But such dogmatism is now untenable: the whole astonishing and para-
doxical character of twentieth century physics shows that our so-called 'cate-
gories, and 'forms of intuition' fail both below and above the limits of our
macroscopic experience preCisely because they are adapted only to what
Reichenbach called "the zone of the middle dimensions". In other words, the
adaptation of the human mind to objective reality is not complete; this
explains both the triumphs of classical physics inside the zone of the middle
dimensions and its failures when we try to extrapolate it beyond its limits.
Consequently, contrary to the author of the Critique of Pure Reason, which
was an epistemological justification of traditional logic, of Euclidean geome-
PARTICLES OR EVENTS? 25

try and Newtonian mechanism, our categories are not a priori nor are they
universally valid; they fit with remarkable precision the zone of the middle
dimensions, but generally fail outside it. Ultimately they are of empirical
origin, and like all other empirical ideas they have only limited applicability.
Now one of the most venerated traditional categories is that of substance.
In the context of our discussion only that of material substance is relevant. It
is certainly striking how little this concept has changed through the centuries,
at least if we disregard the Aristotelian intermezzo which lasted so long only
because of factors foreign to philosophy and science. But in physics and the
philosophy of tlature, the concept of material substance remained essentially
the same. Hume was basically right when he reduced it to a stable conjunc-
tion of sensory qualities, but he failed to explain why these qualities were
restricted to some qualities of sight and touch - while the other, so-called
secondary qualities, were eliminated so early. In other words, using his own
terminology, why the material susbtance, the atom of Democritus, Newton
and Laplace was reduced to a simultaneous conjunction of the basic geometri-
cal and mechanical properties, that is, of mass, space occupancy and motion;
or, in the language of analytical mechanics, to that of position and mo-
mentum. He was apparently little concerned about this problem and this is
why modern classical physicists and philosophers of science generally fol-
lowed Democritus and Locke rather than him, especially when all empirical
evidence then suggested the notion of permanent substance. And not only
empirical evidence: why was the law of conservation of matter anticipated so
early, more than two millennia before its verification by Lavoisier? Kant was
so impressed by it that he regarded substance as an a priori category imposed
on our experience. The point of view of the genetic theory of knowledge is
different: the notion of a material particle quantitatively constant and
persisting through time is formed by the pressure of macroscopic experience
and, as Piaget's research has shown, I think, decisively, in childhood. The
notion of a permanent object, persisting through time, is formed much
earlier, in truth before the end of the eighteenth month, if I remember
correctly. I had a chance to say more about the significance of Piaget's re-
search about three years ago, here at Boston;36 now, I would like to stress or
rather re-stress this: neither the concept of atom, nor that of particle, nor
even that of permanent object are a priori categories of mind, but the results
of our adjustment to a limited range of macroscopic experience; to extrapo-
late them beyond these limits has led, and if we persist in so doing, will lead
to repeated failures. Hence the inadequacy of the concept of a substantial
particle on the microphysical scale;37 this is why the term 'event' is far more
26 MILIC CAPEK

appropriate - it is free of misleading corpuscular associations. What we still


call 'particles' are either individual events, quite often of extremely short
duration or a string of events succeeding each other through sometimes long
intervals of time, but always of finite duration. If I regard 'process', 'event',
'change' as basic categories, it is not because of any subjective preference, but
because they are far more comprehensive and less restricted in their applica-
tions than such terms as 'objects' or 'particles'.

SUMMARY

(a) Oassical physics regarded elementary particles as very minute solid


bodies with constant mass, shape (usually spherical) and volume, persisting
through time and identifiable in the successive positions of their continuous
trajectories.
(b) Not a single one of these features remains unquestioned by contem-
porary physics. The constancy of their mass, shape and volume disappeared
with the special theory of relativity; their permanence with the discovery of
the creation and annihilation of electrons; and the continuity of their trajec-
tories with quantum theory. ..
(c) Even the most abstract residue of this concept - the conjunction of
definite position with a definite momentum - is challenged by wave
mechanics, in particular by Heisenberg's principle of indeterminacy.
(d) Popper-Lan de's defense of particles does not take into account the
strange consequences to which consistent applications of corpuscular models
lead: a return to the original planetary model of the atom, a billiard ball
model of the nucleus, etc.
(e) Fmally, on epistemological grounds, it is exceedingly implausible that
the notion of particle, made in the image of a solid macroscopic body would
apply to the region so remote from our daily experience. The individuality on
the microcosmic level is very probably the individuality of an event rather
than that of a thing.

Boston University

NOTES

1 Cyril Bailey, The Greek Atomists and Epicurus (Oxford: Oxford Univ. Press, 1928);
Kurt Lasswitz, Geschichte der Atomistik vom Mittelalter bis Newton (Hamburg and
Leipzig: Voss, 1890); Emile Meyerson, De /'explication dans les sciences (Paris: Payot,
1921); Federigo Enriques, Le dottrine di Democrito d'Abdera (Bologna: Zanichelli,
1948).
PARTICLES OR EVENTS? 27

2 Cf. an effective refutation of this view in Meyerson, op. cit., II, pp. 320-321, and 356.
3 Benjamin Farrington, The Greek Science (Harmondsworth: Penguin Books, 1944),
Ch.4.
4 K. Lasswitz,op. cit., I, pp. 257-258.
5 Lucretius, De rerum natura, II, vv. 309-333; Bailey, op. cit., p. 332.
6 Bailey,op. cit., p. 80.
7 Enriques,op. cit., Ch. III, '11 principio d'inerzia', pp. 57-91.
8 Hans Witte, Uber den gegenwartigen Stand der Frage nach einer mechanischen
Erkliirung der elektrischen Erscheinungen (Berlin: Ebering, 1906); P. Drude, 'Ober die
Fernwirkungen', Annalen der Physik 62 (1897), pp. XXV-XLIX (on numerous models
of gravitation); finally, E. T. Whittaker, History of the Theories ofAether and Electricity.
The Classical Theories (New York: Philosophical Library, 1951), and Kenneth Schaffner,
Nineteenth·Century Aether Theories (New York: Pergamon Press, 1972).
9 Jean Perrin, Les atomes (Paris: Alean, 1914), p. 253.
10 A. Einstein, 'Relativity and the Ether', in Essays in Science (New York: Philosophical
Library, 1934), p. 110.
11 Cf. White, op cit., pp. 216-219; Osborne Reynolds, The Sub·Mechanics of the
Universe (Cambridge: Cambridge Univ. Press, 1903), p.l.
12 H. Weyl, Was ist Materie? (Berlin: Springer, 1924), p. 18.
13 Cf. the bibliographical references in M. Capek, The Philosophical Impact of Contem·
porary Physics, new paperback ed. (Princeton: Van Nostrand, 1969), p. 242.
14 Niels Bohr, Quantum d'action et noyaux atomiques, Actualites scientifiques et indus-
trielles, No. 807 (Paris, Hermann, 1939), p. 12; Robley D. Evans, The Atomic Nucleus
(New York: McGraw·Hill, 1955), pp. 30-31.
15 Otto R. Frisch, Atomic Physics Today (New York: Basic Books, 1961), pp. 132,
186, and 192; W. Heisenberg, 'The Nature of Elementary Particles', Physics Today 29
(1976), 32-39, "words such as 'divide' or 'consist of' have to a large extent lost their
meaning". Hence Heisenberg's skeptical attitude toward the quark hypothesis (ibid.,
p. 39). (This was probably his last article.)
16 H. Reichenbach, The Direction of Time (Berkeley: Univ. of California Press, 1956),
p.265.
17 E. Meyerson,De I 'explication dans les sciences (Paris: Payot, 1921), I, Ch. X.
18 On this point, see J. B. Stallo, The Concepts and Theories of Modern Physics, Ch. VI:
'The Proposition That All Potential Energy Is in Reality Kinetic' (Cambridge, Mass.:
Harvard University Press, 1960). On Huygens' kinetic model of potential energy, cf.
K. Lasswitz, Geschichte der Atomistik vom Mittelalter bis Newton (Hamburg and
Leipzig: Voss, 1890), II, p. 373. Spencer's view is stated in his First Principles, 4th
edition, Appendix (New York: Appleton, 1896), pp. 598-599.
19 K. Popper, The Logic of Scientific Discovery (London: Hutchinson, 1959), pp.
215 ff.; A. Lande, From Dualism to Unity in Quantum Physics (Cambridge: Cambridge
University Press, 1960).
20 A. Eddington, The Nature of the Physical World (New York: Macmillan, 1933),
p. 225; P. Frank, Philosophy of Science: The Link Between Science and Philosophy
(Englewood Cliffs, N.J.: Prentice·Hall, 1957), pp. 215, 230; Max von Laue, 'Ober
Heisenbergs Ungenauigkeitsbeziehungen und ilIre erkenntnistheoretische Bedeutung',
Naturwissenschaften 22 (1934), 439-441.
21 M. Capek, The Philosophical Impact of Contemporary Physics (Princeton: Van
Nostrand, 1969), Ch. XVI.
28 MILIC CAPEK

22 H. Margenau, The Nature of Physical Reality (New York: McGraw-Hill Co., 1950),
p.313.
23 E. Bauer, 'Rapports entre la physique actuelle et la philosophie', in L 'Evolution de la
physique et la philosophie, Quatrieme Semaine Internationale de Synthese (paris: Aiean,
1935), pp. 31-33.
24 H. Margenau, 'Advantages and Disadvantages of Various Interpretations of Quantum
Theory',Physics Today 7, no. 10 (1954),6-13.
25 E. SchrOdinger, What Is Life and Other Scientific Essays (Garden City: Doubleday,
1958), p. 175.
26 D. Bohm, 'Explanation by Hidden Variables at a Sub-Quantum Level', in Observation
and Interpretation, edited by S. Korner (London: Buttersworth Scientific Publications,
1957), p. 35.
27 J .-P. Vigier, 'The Concept of Probability in the Frame of the Probabilistic and Causal
Interpretation of Quantum Mechanics', ibid., p. 76.
28 P. A. M. Dirac, The Principles of Quantum Mechanics (Oxford: Clarendon Press,
1930), p.4.
29 K. Popper, 'The Propensity Interpretation of the Calculus of Probability and Quan-
tum Theory', Observation and Interpretation, pp. 65-71.
30 A. Lande, op. cit., pp. 78-79.
31 K. Popper, The Logic of Scientific Discovery, p. 221; 'The Propensity Interpretation
... ',p. 69; Lande, op. cit., p. 76.
32 Popper, 'The Propensity Interpretation ... " p. 68.
33 A. N. Whitehead, Science and the Modern World (New York: Macmillan, 1926),
p. 193. There is every indication to support the view that Whitehead was not only fully
aware of the discovery of the vibratory nature of matter, but even anticipated it. The
contrary view of Robert Palter and Abner Shimony is not supported by the texts. Cf. A.
Shimony, 'Quantum Physics and the Philosophy of Whitehead', in Boston Studies in the
Philosophy of Science, Vol. 2 (Dordrecht, Holland: D. Reidel, 1965), p. 307; R. Palter,
Whitehead's Philosophy of Science (Chicago: University of Chicago Press, 1960), p. 218.
On the relation of Bergson's view of matter to that of Whitehead, cf. both my books:
The Philosophical Impact of Contemporary Physics (Princeton: Van Nostrand, 1969),
pp. 368-369, 375 and 391 and Bergson and Modern Physics, Boston Studies in the
Philosophy of Science, Vol. 7 (Dordrecht, Holland: D. Reidel, 1971), Part III, Ch. 14.
34 Whitehead,op. cit., p. 73.
35 Otto von Frisch, op. cit., p. 90.
36 Jean Piaget, Le developpement des quantites physiques chez l'enfant: conservation et
atomisme (Neuchatel: Delachaux et Niestle, 1941); M. Capek, 'The Significance of
Piaget's Research on the Psychogenesis of Atomism', in Boston Studies in the Philosophy
of Science, Vol. 8 (Dordrecht, Holland: D, Reidel, 1971), pp. 446-455.
37 The non-substantia1 character of particles was explicitly stressed by A. March, Die
physikalische Erkenntnis und ihre Grenzen (Braunschweig: Vieweg, 1960), pp. 58-62
and 95-97. W. Yourgrau, although generally favorable to the Popper-Lande view,
concedes that concepts like 'sameness' or 'individuality' do not apply to micro-particles.
See his 'On the Reality of Elementary Particles', in The Critical Approach to Science and
Philosophy, ed. by M. Bunge (New York: The Free Press of Glencoe, 1964), p. 369.
PATRICK A. HEELAN

COMMENTARY ON 'PARTICLES OR EVENTS?'

Professor Mili~ Capek holds that the biological theory of knowledge and
Piaget's genetic epistemology provide an explanation in some sense of the
origin, Significance, and truth value of the atomistic hypothesis. A summary
of his position comprises the following points:
(a) The atomistic hypothesis was conceived neither
(b) as a matter of sheer coincidence nor
(c) because-it is an a priori notion inherent in human reason, but
(d) because it is a notion necessary for human intelligent biological
adaptation to man's middle-sized - neither micro- nor mega- (or cosmologi-
cal) - environment; consequently, the present crises of atomistic theory in
micro- and mega-domains of physics is not a matter for surprise since our
survival to date has not necessitated manipulation either of the micro-domain
or of the mega- (or cosmological) domain.
(e) The 'truth', moreover, of the classical corpuscular-kinetic hypothesis is in
fact that it has contributed a necessary condition to human biological survival.
I shall comment on these points separately.
(a) The atomistic hypothesiS is not just Democritean or Lucretian atomism,
nor is it to be identified exclusively with Newtonian or Daltonian atomism,
nor is it even simply classical corpuscular-kinetic theory, though this last is
the main focus of Capek's examination and critique. The atomistic corpus-
cular or particle hypothesis is in fact what unifies a two-thousand year tradi-
tion of inquiry. It is not exclusively anyone of the theories that are linked
historically by that tradition but an inspiration common to them all - a
common pre-theoretical intentionality that is manifested in all of them and it
is this that constitutes the atomic tradition. It is in the light of this tradition
that a judgment is made as to whether a particular theory is well or poorly
expressed (or formulated), whether a particular theory is or is not in agree-
ment with empirical test-data. It is this tradition that presides over every
change or face-lift that a theory is given as a consequence of confrontation
with empirical data. As Capek has shown, it is not a set of precise and explicit
requirements since there is no requirement that has not been violated in some
one of the historical theories that embodied the tradition.
The historical theories I am speaking about involve mathematical models

29
R. S. Cohen and M. W. Warto/sky (eds.), PhYSical Sciences and History 0/Physics, 29-34.
© 1984 by D. Reidel Publishing Company.
30 PATRICK A. HEELAN

of reality: the truth value of the theory of this kind is in the use of a mathe-
matical model to articulate real structures and real relationships, not merely
in the articulation of the necessary relationships between elements of the
model itself. A scientific model is a construction, an invention, of mathemati-
cal entities: a model does not exist and, despite opinions to the contrary, is
not a substitute for reality. Reality is reached in judgments through the
correct use of model language. No one, for instance, believes or should believe
that the planets or atoms for that matter, are physically non-extended point
particles; yet the classical point particle model has a correct use today as it
had in Newton's time, when speaking of the planets, for example, or of space
vehicles, and sometimes even of atoms. The continuity of a tradition is not in
the series of models used - they may differ from one another as much as a
19th-century steam-buggy differs from an electric automobile of the 21st
century - but in the pre-theoretical intention that animates them and orders
them in a sequence toward the ever more perfect manifestation of the objects
to which they are taken to refer. It is this more or less perfect fulfillment of a
common animating intention in the empirical world that allows- us to order
the theories logically in a series: 1 an ordering that usually is linear and
parallels the historical sequence.
I submit that there is an atomistic tradition in the West which is not to be
identified with classical corpuscular-kinetic theory or with any other parti-
cular form of its embodiment. That atomistic tradition has not been falsified
even by quantum mechanics or relativity; that the crisis that {;apek speaks
about is a crisis of a particular atomistic theory (of the classical corpuscular-
kinetic theory); that this crisis is also a crisis of the atomic tradition but one
of a dialectical sort - and I mean this in a special, technical sense - namely of
conflict between a variety of intentionalities and traditions, each accounting
for wide ranges of phenomena and anticipating, not replacement of one by
another, nor of all by a disparate third, but a synthesis in a new unified tradi-
tion that will account for the ranges of phenomena already accounted for.
I shall return to the notion of dialectical development - so close in some res-
pectsto a biological theory of knowledge, but differing nevertheless from the lat-
ter as Husserl differs from Piaget or intentionality-analysis from psychologism.
(b) The atomistic hypothesis, says {;apek, was not conceived as a matter of
sheer coincidence; hence, there is some minimal Logic of Discovery or at least
of Selection. In question is the discovery of new traditions of scientific
inquiry - not of new empirical facts. What is the origin of a tradition like the
atomistic one? How did it come to have a special significance for scientific
inquiry over such a long historical period? Will it, can it ever be dropped or
COMMENTARY ON 'PARTICLES OR EVENTS?' 31

replaced? Can it be dropped or replaced without a residue? Can we formulate


the minimal laws of discovery and development of scientific traditions?
( c) The atomistic hypothesis, says Capek, is not an a priori notion inherent
in human reason. Thus, he rejects the idealist position: knowledge is not the
projection of meanin~ entirely programmed in and by the individual or
collective subject, a priori to all individual or collective experience. This rejec-
tion is also, in part at least, a rejection of Piaget's position.
(d) The atomistic hypothesis, says Capek, is a notion necessary for human
intelligent biological adaptation to its middle-sized environment (neither
micro- nor mega-); consequently, the present crisis of atomistic theory in
micro- and mega-domains of physics is not a matter for surprise, since our sur-
vival to date has not necessitated manipulation either of the micro- or of the
mega- (or cosmological) domains. Moreover, he adds, the 'truth' of the classi-
cal corpuscular-kinetic hypothesis lies in the fact that it has contributed to
human biological survival.
Capek makes the claim that the atomistic tradition is necessary - not
merely convenient or helpful for human intelligent biological adaptation.
Mere convenience hardly explains, given other alternatives - only necessity
does that. The claim purports to be an attempt at explaining the cultural
persistence of atomistic theories.
But in what sense is atomistic theory necessary for human biological survi-
val? Atomistic theory is necessary for man to manipulate his environment
successfully, so it is said, so as to survive in a possibly threatening, possibly
hostile environment. The notion of manipulation is here taken physically -
to move, touch, grasp, pull, push, put in order, etc. I.et us judge necessity
from two perspectives, from the perspective of a third-party observer, and
from the perspective of the subject's own intention. From the perspective of
a third-party observer, an agent X equipped with the knowledge of classical
corpuscular-kinetic theory can enter into an accomodation-assimilation inter-
action with his environment in such a way as to effect successfully a wide
variety of physical effects, e.g., killing at a distance or travelling to the moon
and back, and by this added power, X is said to have increased his chances of
biological survival for which, consequently, atomistic theory was necessary.
From the perspective, however, of the subject's own intention, he acts with
necessity so as to survive, but for him to survive means to survive humanly.
The sense of the necessitating phrase, 'humanly', escapes all third-party
observers of X's manipulatory or behavioral activities, for to survive humanly
is to incarnate a meaning, to pursue a set of values, to carry forward a tradi-
tion, to be under the burden of moral necessities.
32 PATRICK A. HEELAN

To survive humanly may well include at certain places and times embody-
ing the atomistic tradition as a value. It is the man that accepts this as his
defining human value - the classical scientist, that is, and the culture formed
by the tradition of classical science - that needs the atomistic tradition in
order to survive, that is, to survive humanly in its own way: kill the tradition
and that man and culture lose the sense and meaning of what it is to live -
they lose in some sense their existence.
Since man has begun to manipulate the micro-domain, he has found the
atomistic theory (classical corpuscular-kinetic theory) available to him in-
accurate and unhelpful. Atoms, electrons and elementary particles do not
follow trajectories like classical particles. They are not classical particles. The
breakdown of classical atomistic theory for atomic or sub-atomic systems
should not surprise us, since the classical atomistic tradition was concerned
exclusively with the manipulation of middle-sized entities, entities of the size
we can see, touch and affect in some immediate way. Atoms are 10,000 times
smaller than the smallest thing we can perceive with the naked eye. It ought
not, then, to be surprising to us that new and different physical models are
used. We oUght not to be surprised if the new models are incompatible with
classical corpuscular-kinetic theory: after all, successful manipulation is all
that is required, and whatever succeeds, goes.
But where is one to look for new models? Is there any Logic of Discovery
that sets a priori limits to what we ought reasonably to entertain? Or is the
Darwinian mode - maximum choice and maximum competition - the model
to follow?
Piaget in his genetic epistemology has shown that the knowledge systems
that a child learns in our culture develop according to a pattern - and the
pattern once initiated is necessary in its unfolding. Each knowledge system
represents accommodation and assimilation to a specific group of physical
manipulatory activities. When the child learns a new system, this generally
becomes integrated with the old ones - not by the addition of a new ability
to the old oqes - in a merely arithmetic way - but by the simultaneous
transformation of the old ones so that their efficacy is also broadened and
enhanced. Touching and seeing each has its separate domain: but the domain
of seeing and touching together is larger than the two domains of touching and
seeing taken separately. This mode of integration involves a partial ordering
of systems, arranged within a non-distributive lattice as in Figure 1. 2 Genetic
Epistemology - as a psychological science - says that developing traditions
constitute lattices under a partial ordering of content. Since not all particular
knowledge systems or traditions are, in fact, ever integrated, we should say
COMMENTARY ON 'PARTICLES OR EVENTS?, 33

1= AeB

A B

o
Fig. 1. Non-distributive lattice.

that Piaget's discovery is that the knowledge systems of a child constitute at


any point in time a weak partial lattice of knowledge systems (or traditions).
Such a discovery merely states an empirical fact: it is not normative and as
such has no philosophical import. Let us shift from the psychological or em·
pirical viewpoint of a third·party observer to the philosophical viewpoint of
the subject who is doing the knowing: taking 'philosophical' in a Husserlian
sense, that is, as the attempt to reflect on the normative intentional structures
present and operative in the knower. Then the fact of the peculiar lattice
structure of the development discovered by Piaget appears, I claim, as an epis-
temological norm in the subject: 'Seek to close non-closed lattices of know-
ledge systems'. This principle appears to be one law of a Logic of Discovery. I
call it the Principle of Dialectical Development (or Lattice-Growth): today,
for example, we have S-Matrix versus quantum field theory in a quantum
theory of gravitation.
What I found unsatisfactory in Capek's paper is the implicit promise of a
Logic of Discovery and the absence of any progress towards the enunciation
of what this might turn out to be. Into this gap, I have - immodestly -
thrust some of my own ideas and conclusions. The Principle of Dialectical
Development is one of these. Other principles are usually discovered through
the use of a technique similar to the one I used with Piaget, of combining
empirical historical studies with an intentionality-analysis of the traditions. A
study of the early Einstein and of Heisenberg in the period of his creation of
quantum mechanics reveals five other principles which I shall just enumerate: 3
The Principle of Stable Background Knowledge; The Principle of Model
Continuity; the Principle of E-Observability (Einstein-observability); The
Principle of Contextual Re-Interpretation; The Principle of Pragmatic Con-
tinuity. However, of this set, the Principle of Dialectical Development or
Lattice-Growth is the most important.
34 PATRICK A. HEELAN

(e) Moreover, Capek holds, the 'truth' of the corpuscular-kinetic hypo-


thesis is, in fact, that it has contributed to hwnan biological survival. Hence,
idea-immateriality or idea Platonism is rejected; ideas, he holds, represent
structures of an embodied consciousness, conscience incarnee; mathematical-
physical models consequently for him have merely a technical, instrumental
or manipulatory value; they do not reveal what is in the world. The necessi-
ties of biological survival winnow embryonic scientific theories by a process
of natural selection: success implies truth. Implied in what has been said is
the kernel of a Logic of Discovery; but it is discovery from the perspective of
a third-party observer of the process of discovery, not discovery from the
point of view centered in the successfully inquiring - or discovering - subject.

State University ofNew York at Stony Brook

NOTES

1 The view here expressed is outlined and defended in Part U of my Space-Perception


and the Philosophy of Science (Berkeley and Los Angeles: University of California Press,
(1983).
2 P. Heelan, 'Logic of Framework Transposition',International Philosophical Quarterly
11 (1971), 314-334.
3 P. Heelan, 'Heisenberg and Radical Theoretic Change', Zeit. f Allgemeine Wissen-
schaftstheorie 6 (1975), 113-138, where these principles are enunciated and discussed.
O. COSTA DE BEAUREGARD

TIME SYMMETRY AND INTERPRETATION OF


QUANTUM MECHANICS

1. INTRODUCTION: PARADIGM AND PARADOX

There is something truly paradoxical in quantum mechanics. Physicists know


well how to use it for explaining or predicting an ever-increasing harvest of
phenomena, some of which are striking, even to the expert. Nevertheless, and
notwithstanding the appearance, even recently, of some sedative writings
by both physicists ([1]-[6]) and philosophers of science, it is clear that
many among the leading theorists ([7] -[9]) and the fervent epistemologists
([10]-[13]) remain unsedated by the tranquilizing sort of remedies and keep
suspecting that the problem of properly interpreting quantum mechanics
may well conceal a major enigma.
In other words, the solution of the various well-known paradoxes by
Einstein, SchrOdinger, Wigner, and others may well imply the recognition and
the understanding of one central, major paradox, out of which all the others
are generated.
It seems to me that this problem - understanding quantum mechanics,
and not merely knowing how to use it - has implications somewhat similar to
those of the relativity problem confronting Einstein in 1905. In 1905 the
relativity problem was an old one, originating much further back than the
1887 second-order null effect of Michelson: from the 1818 first-order null
effect of Arago by means of a group-theoretical argument, as emphasized by
Hadamard [4] 1 or even, as ytlmaz [16] convincingly argues, from the 1728
Bradley effect and consistency of Fermat's principle. 2 Through the years it
had been a problem ever more replete with paradoxes, which called forth an
ever-increasing flow of thought, calculations, and experiments. It was a pro-
blem where one knew how to use the right formulas (once these were found:
Fresnel's 'ether drag' formula and the Lorentz-Poincare formulas, already
known to Larmor in 1898 and - not quite exactly - to Voigt in 1887), but
one did not known how to read them. What was missing was just the epis-
temology and the discourse neatly fitting these 'good' formulas and faithfully
rendering, in the world of concepts, their group property. The problem was
thus not one of more sophistication in the mathematics, nor in the axiomatics.
As they stood, in their elegant simplicity, as unveiled by Poincare[17] and

35
R. S. Cohen and M. W. Wartofsky (eds.), Physical Sciences and History of Physics, 35-56.
© 1984 by D. Reidel Publishing Company.
36 O. COSTA DE BEAUREGARD

later by MinkowskV the mathematics was all right. Playing with the axioma-
tics, as has been long fashionable after Einstein's discovery, was not the
answer either. No, the problem was just plain reading of the Lorentz-Poincare
formulas or of the Fresnel ether drag formula([14], [15]), faithfUlly ren-
dering their group property. This would bring in quite smoothly the relativity
of time and space, however paradoxical this epistemology has seemed to be
and this discourse has sounded. The prophet, of course, who unveiled the
sense of the scriptures was Einstein, in his 1905 paper, where none of the
mathematics is new. The breakthrough lay entirely in the interpretation, thus
bringing all the old paradoxes to the point of radiance as one dazzling, but
illuminating, new paradox - very much as Copernicus had done in older days
and under other circumstances.
This bring; me naturally to the crucial role of paradoxes and their akinness
to paradigms [18]. One reads in Funk and Wagnall's Standard Dictionary of
the English Language the follOwing definition: "Paradox: 1. A statement,
doctrine or expression seemingly absurd or contradictory to common notions
or to what would naturally be believed, but in fact really true." There is no
doubt that Copernicus and Einstein's statements and doctrines have been
paradoxical in this primary sense, and that they exemplify a process that is
quite common in the advancement of science. Science, in its acrobatic
advance along the rope, always oscillates between modelism and formalism.
Maxwell's and Boltzmann's statistical mechanics (but not Gibbs') can be
taken as a victory of modelism, while Kepler's three laws, or Einstein's and
Minkowski's relativity, are triumphs of formalism. Formalism, in its clever
Simplicity, dissolves the clumsy constructs of modelism - for instance, the
mechanical theories of the ether - very much like that stone in the Book of
Daniel (Bin Stein), which came full speed from elsewhere, and reduced the
composite colossus to pieces.
It also happens that the new synthesizing paradox assumes and gives sense
to a few small paradoxes (in Funk and Wagnall's fundamental sense) which
have kept creeping in through the ages and were taken as superstitions. For
example, falling meteorites, an obvious fact to farmers or hunters, were still
a superstition to Laplace in the eighteenth century, while in the early seven-
teenth century a scholar wrote that "Briton sailors are so superstitious that
they believe the Moon has an influence upon the tides." One need not say
that these two superstitions have become part of the religion in Newton's
gravitation theory. The paradox of action at a distance has assumed both of
them.
Finally, the position I am taking is that the problem of interpreting or
TIME SYMMETRY AND QUANTUM MECHANICS 37

understanding quantum mechanics is essentially one of plain unbiased


reading of the 'good' (that is, the very operation and elegant) formalism we
already have, thanks to Heisenberg, Schrodinger, Dirac, and others. This I
believe - and intend to show - will uncover a very new but also illuminating
central paradox, one that will incidentally assume and perhaps render more
respectable one or two superstitions.

2. COLLAPSE OF THE WAVE PACKET: AN ACTIVE INTERVENTION


OF THE PSYCHE

Among the ever-flowing deluge of papers devoted to the· interpretation of


quantum mechanics, I select Moldauer's ([3], [4]) very valuable work, not
only because it aptly clarifies some important technical points, and it is
representative of those ([1] -[5]) who, by asking, "Is there a quantum
measurement problem?" imply that there is none, but mainly because it
shows quite clearly at which point the central issue is sidestepped. Moldauer
writes [3]: "This paper does not deal with ... the following questions: Is a
probabilistic theory a sufficient description of phenomena that have only
statistically reproducible properties? What is the precise meaning of proba-
bility?" These are the embarrassing issues I am aiming at. As long as they are
swept under the rug, the dust will keep trickling out, and then the most clever
axiomatics or competent mathematics will not prevent the audience from
coughing from time to time. This I do, when Moldauer writes elsewhere [3],
"The consequent reduction of the state vector ... is completely described by
the combined tools of the Schrodinger equation and the statistical interpre-
tation .... "
The critical point is the interpretation of the stochastic event that occurs
when a transition takes place, and when one among the various a priori
possibilities (each endowed with a suitable probability) is actualized. In
classical statistical mechanics it was assumed that stochastic events do occur,
the frequencies of which reproduce (in repetitive tests) the calculated proba-
bilities. Mutatis mutandis (and, of course, some of the mutata are highly
specific), this classical trait reappears in von Neumann's quantal ensembles.
But this is definitely not the point, and various axiomatic approaches
expertly playing with the density matrix sidestep it.
The central problem is what occurs in the individual quantal transition -
the preliminary answer being that nothing in the quantal formalism tells us
what. I believe I have read in an article (but know not where, so perhaps I
have dreamt it), that there is nothing in the quantal formalism speaking
38 O. COSTA DE BEAUREGARD

by itself of an event. 4 This is the point, and it is very clearly stated in two
early works: von Neumann's [19], where it is implicit everywhere, and
London and Bauer's [20], where it is quite explicit. And the very same
answer to the problem was put forward in these works: The event or transi-
tion that is expressed formally as collapse of the state vector (into one of its
orthogonal components) occurs when, and only when, the observer takes
cognizance of the experimental result. Thus, the quantal stochastic event is
neither purely objective, because it would not occur in the absence of some
sort of consciousness registering it, nor purely subjective, because it truly
occurs in the real world. In other words, the quantal stochastic event must be
thought of as indissolubly objective and subjective - a trait which I have long
believed [21] -[23]) to be intrinsic to true or essential probability - if only
because the purely objective and the purely subjective schools both run into
severe difficulties. Thus it may well be that the quantal stochastic formalism
is much nearer than the classical one to an adequate expression of essential
probability. In this respect Lande [24] has Significant things to say, but I will
not delve into them here.
Perhaps I should quote, as supporting what I am saying, a few sentences
from a very searching article by Hooker [12]. Hooker argues against both
Jauch's treatment of the Einstein-Podolsky-Rosen (EPR) paradox and Krips's
treatment of the Schrodinger cat (SC) paradox. He writes, "One wants to
know what precisely is physically going on in a single given instance of the
.measurement process when this transition is supposed to be occurring. No
answer seems to be forthcoming from quantum theory." And a little later,

Which one of the statistical possibilities is in fact realized when the measurement pro-
cess is over is not represented in the theory until some human observer 'takes a look' and
decides on the basis of that look to change the state representation from the statistical
mixture to some particular pure state. This kind of change is commonplace, of course, in
classical statistical theories, and it provokes no comment there precisely because we do
not take them to be offering a complete description of physical reality. But only
Einstein and the like-minded have continued to argue this status of quantum theory
itself.

This being said, I certainly do not pretend that Prof. Hooker is ready to
follow me in what I will be stating later.
Concluding this section, I cannot see any possible escape from the twin
statements: (1) Quantal transitions, as a specific sort of stochastic event,
do occur, and they imply a discontinuous jump, or collapse, of the state
vector; and (2) as von Neumann, and London and Bauer, have stressed,
TIME SYMMETRY AND QUANTUM MECHANICS 39

the state vector collapse in an individual transition is due to an act of con-


sciousness on the observer's part.

3. INTRINSIC STATISTICAL TIME SYMMETRY AND ARISTOTLE'S


TWOFOLD INFORMATION CONCEPT

Let me recall that the problem of understanding physical irreversibility in


terms of probability theory has turned out to be far more subtle than it had
seemed at first. The appropriate answer to the well-known Loschmidt and
Zermelo paradoxes has been clarified only recently by quite a few physicists
and/or philosophers of science, who had started thinking independently and
have come to an essential agreement, even if they stress different aspects of
the question, or perhaps differ on some minor points. 5
&sentially, the irreversibility paradox is inherent at the very start of
probability theory, where it is given a technical answer by means of Bayes'
conditional probability formula. The paradox is that, given some initial
complexion of a stochastic system - say, a deck of cards - a 'blind statistical
prediction' of the issue of a test, or transition - say, of shuffling the cards -
will be operational, while a 'blind statistical retrodiction' will not.
Nobody will rely on shuffling to put the deck in order. At least this is the
situation prevailing in physics, where of course it is very tightly connected
with the second law of thermodynamics. A radioactive nucleus will decay
spontaneously according to the laws of probability, but no physicist expects
that stray electrons and neutrinos will converge toward a cell containing 11 B
and fill it with 11 C. In other words, blind statistical prediction is physical
while blind statistical retrodiction is not. And this cannot be accounted for
by the intrinsic transition laws, which are, in most cases, taken to be time
symmetric6 - as, for instance, in card shuffling or in radioactivity.7 For this
reason, retrodictive physical problems are treated with the aid of Bayes's
conditional probability formula, that is, by using not only the intrinsic
transition probabilities, but also a set of extrinsic probabilities (Bayes'
coefficients), which are estimated according to one's idea of the situation,
and obviously represent at best the initial interaction out of which the
stochastic system under consideration is born. This use of Bayes' formula
is the technical trick expressing the fact that blind retrodicition is forbidden
in physics. To my knowledge, van der Waals [25] was the first to point out
that the statistical interpretation of Carnot's principle is a specification of
Bayes's prinCiple, while the same idea is implicit in an often-quoted sentence
by Gibbs [26].
40 O. COSTA DE BEAUREGARD

But knowing how to express the fact that blind statistical retrodiction is
forbidden does not explain why physical interactions macroscopically
produce after-effects rather than before-effects. If, between times t1 and t 2 ,
a physicist moves a piston along the wall of a vessel containing a gas in
equilibrium, Maxwell's velocity law will be disturbed after time t 2 , but not
before time t1. The perturbation will be emitted as a divergent, or retarded,
pressure wave and not absorbed as a convergent, or advanced, wave. This
example displays a one-to-one connection between the principle of increasing
probability - the second law - and the principle of retarded waves - the
principle of causality. This point has been fully clarified only recently,S
but the classical physicists must have guessed it in some sense, since they
termed their use of Bayes' principle in retrodictive problems the principle
of probability of causes. Symmetrically, there is also a one-to-one association
between the two (unphysical) principles of decreasing probabilities and of
advanced actions; on this side, the connection with the philosophical
concept of finality has always seemed obvious. let me mention Bergson [27]
as making a strong case of finality as an (at least seemingly) anti-Camot
process, and the Italian mathematician Fantappie [28] as conceiving finality
as an advanced wave process.
So, the search for the root of physical irreversibility leads to the con-
clusion that it is not at all intrinsic in the elementary laws of evolution
(see Note 8), but rather that it emerges macroscopically as a boundary
condition imposed upon the integration of the Boltzmann or the 'master'
equation. This is in striking analogy with the physics of waves, where
irreversibility is similarly absent from the so-called wave equation, and
appears only via the boundary condition chosen when the equation is
integrated. This suggests the existence of a physical connection between the
two statements, which can indeed be displayed in the realm of quantized
waves, as I will discuss in the next section.
Now, as I have said, my philosophy in this paper is to rely completely
on the formalism, so that interpreting the formalism builds an epistemology
isomorphic to the intrinsic symmetries of the formalism, just as a well-cut
dress is isomorphic to the body. This it seems to me was Einstein's work as
founder of the relativity theory. Or, to take an example better suited to our
symmetry problem, the 'hole' in Dirac's electron theory has been exactly
filled by Anderson's positron - a quite unexpected and rather rare phenom-
enon. The de facto very large dissymmetry between the rare positron and the
trivial electron does not preclude their de jure complete symmetry. I thus feel
logically justified in taking anti-Camot processes, that is, advanced action
TIME SYMMETRY AND QUANTUM MECHANICS 41

processes, as macroscopic ones that are not strictly forbidden, but are usually
rare, or at least do not occur in the typical physical context. It then remains
to be discussed whether perhaps they could not appear under appropriate con-
ditions. If so, and if the twin (macroscopic) principles of increasing probability
and of retarded actions are taken to be an essential part of physics, then the
hypothetical context I am alluding to should be termed antiphysical - very
much as the positron is termed an antielectron. I am well aware that this
direction leads me straight toward paradoxes in the strongest sense. This
I will accept boldly, remembering not only the dictionary's defInition, but
also that the 'strange' world of antiparticles has become a scientific EI
Dorado. Let us run the risk that perhaps the very strange world of anti-
physics might also be a scientific EI Dorado....
The intrinsic time symmetry I am discussing has important consequences
in the question of physical equivalence between negentropy and information.
The discovery of this equivalence, which is the very heart of cybernetics,
is another instance of multiple independent discovery by mathematicians,
physicists, and engineers. 9 Let us equate the essence of this discovery with
the two faces of a medal: heads and tails. Are we not speaking of games of
chance?
The first major discovery of cybernetics, as Gabor put it, is that "one
cannot get anything from nothing, not even an observation". One cannot
obtain information by reading, listening, or sensing in any way, without the
negentropy of the environment diminishing by an amount at least equal to
the information that is gained. Both concepts, information and negentropy,
are defined through the same mathematical formula: the logarithm of a
probability. Thus cybernetics interprets the gain of knowledge - "getting
information," in the words of the man in the street, when for instance he
buys a newspaper - as a generalization of the passive Carnot process.
Instead of letting the negentropy of a closed system become uselessly
degraded, one can recapture part - or, ideally, the whole - of it, in the form
of knowledge.
The other facet of the discovery is that existing information can be used
to produce macroscopic order, the 'negentropy' thus generated being at
most equal to the information that has been invested. A typical instance
of this is the activity of Maxwell's demon, as intrepreted by Brillouin and
other cyberneticists. In this respect information appears as an organizing
power or, in other words, as power of action or of will.
What is truly astounding is that for Aristotle - the proponent of both
the concept and the word - information was a towfold entity: knowledge
42 O. COSTA DE BEAUREGARD

one could acquire, and an organizing power one could use. Without having
sought it, cybernetics has precisely hit upon the two facets of the Aristotelian
concept. .
Now, the fact (if not the legal right) is that the first Aristotelian meaning,
gain of knowledge, is trivial to everybody, while the second one, organizing
power, is somewhat esoteric and familiar only to those few philosophers
interested in will and fmality. I believe ([21]-[23]) this fact to be a mere
corollary of the other 'tact noted above, the extreme preponderance of
entropy- (or probability-) increasing processes lO over decreasing ones, which
is equivalent, as we have seen, to the preponderance of retarded (macro-
scopic) waves over advanced ones. This implies as a consequence an extreme
preponderance of the passive, learning transition, over the active, willing
ones. Very much like positrons among the crowd of electrons - needles
in a haystack - so are advanced action phenomena scattered among the
Niagara cascade of retarded action phenomena. That is, so are final processes
as compared to causal ones. Or (in terms of the subjective side of the
probability concept), conversely, so is willing awareness more strongly
sensed than learning awareness.
It should be obvious that the very values of the universal constants of
physics in terms of 'practical' units directly reflect man's existential situation.
For instance, if Einstein's constant c, the speed of light, is very large when
expressed in, say, meters and seconds, it is because man finds it convenient
to view meters and seconds as associated standards of length and time. This
may very well be because the velocity of our nerve impulses is of the order of
meters per second. For this simple reason the relativistic phenomena lie far
beyond the observation range of everyday experience. Things would be com-
pletely different if the velocity of our nerve impulses were some large fraction
of c. Mutatis mutandis, I believe the situation to be very much the same with
respect to negentropy and information. The conversion coefficient between a
negentropy expressed in 'practical' thermodyn3mic units and an information
item expressed in its natural binary unit is Boltzmann's constant k (more
precisely, it is k In 2), and this is quite small. Thus, gaining knowledge is
extremely cheap in negentropy terms, while producing negentropy costs a lot
in information terms. This existential state of affairs directly reflects the fact
that our world is a Camot world, where retarded actions outweigh advanced
actions.
Significantly, I believe, the universal constants of the major twentieth
century theories are exceedingly small, or large, as expressed in 'practical'
units. Besides Einstein's c and Boltzmann's k,11 the other example is of
TIME SYMMETRY AND QUANTUM MECHANICS 43

course Planck's h. The implication is that all these important aspects of twen-
tieth century physics lie far outside the domain of man's everyday experience.
Now, it is a familiar sort of exercise to see how, by taking an extremely
small universal constant to be zero (or a very large one to be infinite), one
falls back on the familiar state of affairs and loses the far-reaching, or
'paradoxical,' insight that comes with scientific novelty. Thus, by taking
lIe to be zero, one loses Einstein and recovers Newton, or, by taking Planck's
h to be zero, one loses Einstein's photons in optics and de Broglie's matter
waves in mechanics.
What are we losing if we take Boltzmann's k to be zero? We render
learning more than cheap: gratuitous. And we render acting through will
more than costly: impossible. This is a theory that was very fashionable
in the nineteenth century under the name of epiphenomenal consciousness.
The cybernetic discovery is that consciousness, as a spectator, must
buy its ticket for one dime or two. But this alone is sufficient for allowing it
to become an actor also. Thus, our task is now to look beyond the de facto
Carnot situation that hides the deeper questions by properly 'shuffling the
cards'. We have to understand the de jure symmetry concealed behind the
de facto asymmetry. And this might well expose novelties more paradoxical
than the positron.
A quotation from Brillouin [30] may be in order at this point. He writes:
"Relativity theory seemed, at the beginning, to yield only very small correc-
tions to classical mechanics. New applications to nuclear energy now prove
the fundamental importance of the mass-energy relation. We may also hope
that the entropy-information connection will, sooner or later, come into the
foreground, and that we will discover where to use it to its full value."

4. INTRINSIC QUANTAL TIME SYMMETRY AND COLLAPSE OF


THE STATE FUNCTION

Although the specific rules of the original sort of probability calculus


inherent in quantum mechanics are markedly different from the classical
ones, all the essentials of the preceding analysis are retained in it. Discarding
the rare case of some weak interactions that are T-violating, we can state
that the intrinsic predictive probability that a quantum state A goes into a
quantum state B is equal to the intrinsic retrodictive probability that state B
has come from state A. This is the intrinsic stochastic time symmetry of
quantum mechanics, which must be reconciled with the macroscopic time
dissymmetry so evidently displayed, for example, in radioactive decays or
44 O. COSTA DE BEAUREGARD

electromagnetic radiation. The technical answer is the same as before:


Whenever we are speaking not of one single quantal transition, but of a
macroscopic ensemble of transitions, blind statistical prediction is physical
while blind statistical retrodiction is not. And here, in quantum mechanics,
the one-to-one association of increasing probability and retarded waves on
the one hand, and of decreasing probability and advanced waves on the other
hand, is very much tighter than in classical statistical mechanics. This is
because the probability concept and the wave concept are both inherent in
quantum mechanics - an association that might well be of very deep signifi-
cance for natural philosophy. As Fock [31], it seems, was the first to state
explicitly, retarded waves are used in quantum mechanics for statistical
prediction, while ~dvanced waves should be used for statistical retrodiction.
Thus, 'prohibition of blind statistical retrodiction' and 'macroscopic
nonexistence of advanced waves' are merely different wordings of the same
statement. Incidentally, this reconciles the apparently contradictory opinions
of Einstein and Ritz in their famous controversy [32]. For Ritz, probability
increase implied a postulate of wave retardation, while for Einstein wave
retardation should be understood as a consequence of probability increase.
It is clear today that these two statements are reciprocal. Moreover, both
express a de facto situation expressible as an initial (not final) boundary
condition. If this was not clear in the days of Einstein and Ritz, it is because
while Einstein's light quanta were then known, de Broglie's waves were not.
It was thus not obvious that wave scattering and particle scattering go
physically hand in hand.
A very neat way of displaying ([21] -[23]) the reciprocity of the two
laws of probability increase and of wave retardation uses the formalism of
von Neumann's [19] quantal ensembles. It is, in fact, a mere rewording of his
irreversibility proof. The quantal entropy is found to increase at each
transition if retarded waves are used between two interactions. Otherwise,
the entropy would decrease.
The interpretation of quantal entropy as information is quite transparent
in von Neumann's well-known book. The explicit demonstration has been
given by Jaynes [33] in the second of his two pioneering articles on statistical
mechanics. It had been partially expressed earlier by Elsasser [34]. In fact,
the quantal version of Jaynes' reasoning is a truly natural sequel to von
Neumann's quantal ensemble theory.
With Jaynes' formalism at hand, it is merely routine to extend, in
quantum mechanics, all that has been previously said of Aristotle's twofold
information concept, and of the intrinsic symmetry between gain in know-
TIME SYMMETRY AND QUANTUM MECHANICS 45

ledge and organizing power. So, here again, we face the problem of giving an
operational meaning to that semi population of entities that seem to exist
only as abstract concepts: macroscopic advanced waves and information as
organizing power. However, we have in our hands a magic wand for con-
ferring life to these ghosts, a magic wand that was missing in classical statisti-
cal mechanics. As stated by von Neumann [19], London and Bauer [20],
Wigner [9], and others, the quantum event occurs if, and only if, there is an
active intervention of the psyche. So now we really have to understand what
sort of being is the (still evanescent) ghost appearing in our formalism. We
have to track our positron.
One last remark is in order, however, and it pertains to relativistic
covariance. Relativistic covariance and waves naturally belong together,
as explained by Einstein in 1905 and by de Broglie in 1925. The basic wave
equations of quantum mechanics are all relativistically covariant, and
Tomonaga, Schwinger, and Feynman have endowed the quantum field
theory with complete relativistic covariance. What is perhaps less well known,
and is at present important, is that the basic, so·called 'first quantized'
formalism lends itself very well to full relativistic covariance.
Following a hint by Riesz [35], I have developed this formalism in articles,
and fmally in a book [36], while Wightman and Schweber [37] were pro-
ducing similar formulas. Thus there will be no problem with relativistic
covariance when we later tackle the EPR paradox (among others).
But relativistic covariance entails a more fundamental lesson. It has been
said that relativity theory had lost the subject of the verb to undulate. If so,
wave mechanics or quantum mechanics has hit upon the unforeseen subject
of the verb - and one very different indeed from the lost ether. What is
undulating through the vacuum, as explained by Dirac [38] and by Lande
[24], is the amplitude of the probability. Dropping technical precision,
we can speak of quantal waves as probability waves, or information waves.
That is, when we speak of von Neumann's ensembles, retarded, predictive
waves are waves of cognizance, and advanced, retrodictive waves are waves
of will. 12

5. BRINGING A GHOST TO LIFE AND SCHRODINGER'S CAT


PARADOX

The ghost present in the formalism, thought to be absent from reality, is


advanced waves in some macroscopic sense, displaying an existence of
finality, or, in other words, an operational character of willing awarenesS.
46 O. COSTA DE BEAUREGARD

Where are we to look for this anti-Carnot, anti-causal, anti-learning forlorn


twin? The symmetry principle guiding us is not the same one as that behind
the search for antiparticles, but our hope rests on a similar faith - that
symmetry in the formalism is not misleading.
First we must discuss the stochastic event expressed as collapse of the
state vector; it is also termed quantal transition. If this collapse needs the
active intervention of some psyche, even in the cognitive or 'passive' case of
impartial observation', it precludes independence of two observations of the
same event. They must cooperate (or perhaps compete) in producing the
result. And if they occur in succession, it seems that usually (but not
necessarily always) the first one has far more weight, and that any following
one is then bound to confirm its finding. One need not say that these
statements imply a very drastic reinterpretation of von Neumann's arbitrary
severance between the observer and the observed system.
For instance, in the Schr6dinger cat problem there is by hypothesis a
first informed observer - the catY The psyche (if any) that produces the
wave collapse is the cat's - and usually not that of any of the psychobiolo-
gists versed in quantum mechanics who theorize before opening the box.
But if we believe in symmetry between cognizance and will, we are
logically led to the working hypothesis that collapse of the state vector
can be caused not only by knowing awareness, but also by willing awareness.
If so, the cat should be able to influence the yes-or-no outcome to which he is
subjected. And if so, one guesses that a normal cat will be in favor of the yes.
At this very point we are hitting upon a 'superstition', upon a paradox
from below, like that of meteorites, or of the Moon causing tides. For what
we are speaking of has a name in the realm of parapsychology, and that name
is psychokinesis.
Since a paradox, even one of the below kind, can certainly not be refuted
by simply stating that it is contrary to common sense, we are led to inquire
if perhaps parapsychologists have not already performed the Schrodinger
cat experiment as we have just defined it. Well, they have; and if not with
cats, at least with rats or with cockroaches. And if not as a 'death or life'
dilemma, at least in a 'reward or punishment' fashion. And, as I have been
told, they have done it with consistent success. They have found that the
statistics of a random-outcome generator, tested before and after the psycho-
kinetic experiment (and found to be perfectly normal in both tests), are
systematically deranged when the animal is in the box, and this, of course,
in the way favoring the animal - that is, more rewards, or less pUnishments.
This kind of experiment has been done not only with classical, but also with
TIME SYMMETRY AND QUANTUM MECHANICS 47

quantal random outcome generators. The latter case is of course by far the
most directly Significant. The experimentalist is the physicist, Schmidt ([39],
[40)).
Perhaps I should also quote a letter to the Editor of Science [41]. It reads:
"During the past year I have had some correspondence with J. B. Rhine
which has convinced me that I was highly unfair to him in what I said in an
article published in Science in 1955 (26 Aug., p. 359). The article discussed
possible fraud in extrasensory perception experiments. I suspect that I was
similarly unfair in what I said about S. G. Soal in this paper. Signed:
George R Price".
So let us proceed. It is trivial to everybody that a single statistical
quantal outcome, say, that an electron from a decaying radionuclide goes
or does not go through a Geiger counter, with respective probabilities n/41T
o~ 1 - n/41T, is recorded via an amplifying procedure using macroscopic
retarded waves (perhaps in the form of the ultimate feelings of a cat). But,
symmetrically, it sounds fantastic that an animal inside a box, where he is the
innocent toy of a reward-or-punishment stochastic gadget (working through
quantum statistics), can learn enough about what is going on that, by looking
backward in time through the amplifying mechanism (whatever it is) by
means of advanced waves, he should be able to act upon the elementary state
vector collapse so that this transition, instead of being the source of a
retarded wave, as usual, is the sink of an advanced wave. This is 'paradoxical'.
But it is logical, as soon as we believe that matter waves are information
waves, and that all their stochastic formalism is intrinsically time-symmetric.
If we then call reading the (causal) use of retarded waves, we should term
antireading the (final) use of advanced waves.
One need not emphasize that the taboo we here trespass against was
labeled 'no reaction to the observer's glance by the measured system'.
This taboo should be taken as de facto rather than de jure - and we are here
deliberately taking liberties with good manners. Also, it has often been
written that it is the finiteness of Planck's constant which obliges one to
consider the reaction of the measuring device upon the measured system.
Things are not exactly so. What the finiteness of Planck's h makes real is the
one-to-one binding between increasing entropy and retarded waves. The
reaction of the observer's glance upon the measured system is brought in by
the finiteness of Boltzmann's k, and was already inherent in the very concept
of Aristotle's twofold information. In other words, it is inherent in the very
idea that the probability concept is both objective and subjective, being the
hinge around which matter and psyche are interacting.
48 O. COSTA DE BEAUREGARD

6. EINSTEIN-PODOLSKY-ROSEN PARADOX

The Einstein-Podolsky-Rosen paradox [42] is perhaps the most famous


of the quantum paradoxes. It is a variant of other paradoxes due to
Einstein [43], Schrodinger [44], and Renninger [45]. I have more than once
thought (or dreamed) about the vicious sting intrinsic in the EPR paradox,
and it is this meditation which, more than anything else, has convinced
me 14 that the essential nonlocality displayed in it (and analyzed by Bohm
[49], Bell [50], Shimony [51], and others) cannot be understood othelWise
than through the intrinsic past-future and cognizance-will symmetries.
The first thing t\lat should be made clear is that the distant correlation
in the EPR paradox is defmitely not of the trivial sort existing in classical
statistical mechanics. If, say, a positronium atom rotating at point r = 0
is made to explode at time t = 0 by an energy (and momentum-less) exci-
tation, a measurement showing that the decay electron passes through point
rA at time tA = rAv makes sure that the decay positron passes at time
tB = tA through the distant point rB = -rA. This is because of the law of
conservation of momentum and has nothing paradoxical if one believes in
hidden determinism. The quantal sort of correlation is of a more subtle
character, as emphasized by d'Espagnat ([7], pp. 99-139) and, of course,
by the calculations of Bell and Shimony. There is, nevertheless, a lesson
to be gained ([46] -[48]) from the classical case. It is that the logical
inference that deduces what is happening around the spatially distant point-
instant (rB. tB) from the measurement made around the point-instant
(rA, tA) is not telegraphed along the spacelike vector (rB - rA. tB - tA = 0),
but rather along the two timelike vectors (-r, - tA) and then (rB, tb). Of
course the symmetric statement would hold if an inference were drawn as to
what happens in (rA. tA) from a measurement in (rB. tB)' Also, nothing, of
course, prevents two observers 0: and (3 from operating one on the electron A
and the other on the positron B. That sort of space-time telegraph works
both ways.
Why then does the quantal statistical correlation have very surprising
characteristics that are absent in the classical one? Because, of course, the
mathematics is different. More specifically, Bell and Shimony have shown
that the form of the mathematics of the quantal correlation in the EPR
situation is incompatible with the idea that the two diverging subsystems
are governed by some hidden parameters belonging separately to each of
them. 15 Therefore we have what d'Espagnat calls the nonseparability of these
two subsystems (which, by hypotheSiS, have interacted in the past or, in the
TIME SYMMETRY AND QUANTUM MECHANICS 49

Minkowski space-time scheme, are indeed coupled sub specie aetemitatis).


Also, in an analysis by d'Espagnat, it is emphasized that if the two subsystems
were thought to have each their own attributes, their union would be
represented as a (particular) mixture. What are, then, the quantal facts?
At any time the total evolving system can be represented as being
potentially a mixture with respect to the orthogonal states corresponding to
a pair of associated possible measurements, but then not as a mixture with
respect to the states of a pair of associated magnitudes that lire not simul-
taneously measurable with the preceding ones. We are speaking, for instance,
of the pair of x spin components and of the pair of y spin components of the
electron and positron issuing from a spin-zero positronium atom. Until
the very last moment, observer cx, say, may hesitate as to which Cartesian
component of the spin of particle A he will measure. However, as soon as
he makes up his mind and performs the measurement, he is sure that if
observer {3 measures the corresponding spin component of B, he does find
the value that is strictly correlated to the one he has found himself. In this
case we have not only telediction, as in the classical situation, but also
teleaction, in the sense that, when cx performs his measurement on A, a transi-
tion truly occurs there, and that the same transition certainly occurs in B if {3
performs the measurement corresponding to that of cx.
Now, the formalism clearly shows that this telediction-and-teleaction is
telegraphed not directly along the spacelike AB vector but (like the teledic-
tion in the classical case) along the two time like vectors AO and OB (O being
inside the space-time domain where the two subsystems are generated). The
AOB or BOA zigzag is similar in many respects to a Feynman zigzag.
In my philosophy, where advanced actions are postulated to exist, and
to be operational in some specific cases designed ad hoc, the EPR situation
is taken to be one of these. My philosophy thus escapes the ritual EPR
sentence, "if, without in any way disturbing a system, we can predict with
certainty ...." First, since the AB vector is taken to be spacelike, I must
replace the word predict by teledict. And second, the telediction is also a
teleaction - with the relaying satellite placed in the past.
Before explaining how relativistic covariance is preserved in all this I
should speak of experimental verifications of my idea.
First, one could take as cx and {3 two impartial, passive, observers. In
this form the EPR experiment has been done many times, including those
recently inspired by Shimony ([51] -[53]).
Second, one could take an impartial observer as {3 and a selected or
trained psychokinetic agent as cx. This would be an experiment in psycho-
50 O. COSTA DE BEAUREGARD

kinesis, with the observer looking not over the agent's shoulder (as usual),
but along a lateral channel. It would certainly be an interesting experiment if
performed with a sophisticated apparatus of the Shimony family.
Finally, one can take two trained psychokinetic agents as a: and (3, and
have them either compete or cooperate. This would also be an interesting
experiment.
As with relativistic covariance, there is 'no problem'. The measurements
by a: and (3 are both performed inside limited space-time domains, which can
be thought of as extremely small with respect to the spacelike distance
rA - rb and to the time distance tA = tB. In fact we are working with
propagators or relativistic Green's functions, attached to the two vectors
OA and OB. Relativistic covariance is obvious.

7. WIGNER'S FRIEND

And what if two observers a: and (3 look at the same recording apparatus
0, which we take with Wigner [9] to be quantal, that is, not macroscopic
in the sense of Ludwig [5], or of the Prosperi [54] group? The recorded
measurement is transmitted, via information waves (saY,electromagnetic
waves) between 0 and both 0: and (3. And, by the very hypothesis, both a: and
(3 are collapsing, strictly coupled states in the EPR sense. Thus what we have
is akin to the EPR situation.
And what if we follow Wigner and insist that somebody, for instance 0:,
describes the total system also after the measurement by (3 has been made?
Moldauer ([3], [4]) has thoroughly discussed the technicalities of this
problem. Operationally speaking, it is hard to conceive what sort of
measuring apparatus would be able to test the phenomenology of the com-
bined system. Looked at philosophically, however, the question makes sense
and raises as a following question that of a hierarchy of superminds looking
over each other's shoulders.
I will not delve into this near-to-metaphysical problem, but rather fall
back on phenomenology and feasible experimentation. Observer (3, after Ills
measurement, is certainly no more in a linear superposition of states, until
a: finds out which is which, than Schrodinger's cat is before the biologist
opens the box. Here, again, what we have is competition or cooperation
between active psyches who are producing the state vector. In the Schrodin-
ger problem it seems that the cat is more strongly motivated, and less
indirectly coupled, to the decaying atom, so that his decision has a priori far
more weight than that of the biologist. However, I do not exclude that some
TIME SYMMETRY AND QUANTUM MECHANICS 51

sort of telepathic experimentation between the cat and the biologist could
make sense in Schrodinger's context. On the other hand, in the EPR context,
the very symmetry postulated between the a and (3 (real or virtual) observers
has led us to conceive a 'fair contest' between a and (3, as if (mutatis mutandis)
they were pulling the ends of a rope going over a pulley situated in the past.
It seems to me that Wigner's problem lends itself to any specification between
these two extreme cases. One thing is certain however, and Wigner states it
quite clearly: No observer (neither cat nor experimentalist) can have his
mind in a superposition of states, because it is an act (either cognizance or
will) of that mind that collapses the state vector.
Now I am well aware that this leaves me with a very serious problem I
must fmally discuss.

8. WHAT IF THERE IS NO CAT IN SCHRODINGER'S BOX?

Let us recall the situation. We have ideally, say, one single (3-radioactive
atom enclosed in a little cell around point r = 0 and time t = O. Its half-life
is much smaller than, say, T, so that, reasoning predictively, we feel confident
that when we open the box at time T the atom has decayed. Now, the (3
electron may either trigger a Geiger counter seen through the solid angle n
from point 0, with a priori probability n/41r, or else pass beside the Geiger
counter with a priori probability 1 - n/41r. When triggering the counter,
the quanta! event induces a cascade and thus a macroscopic event in the
sense of classical physics. However, the hypothesis now is that the recording
apparatus is no longer a cat, but merely any physical recorder you like.
Nobody on earth, including von Neumann or London and Bauer, would
have it that the die is cast at the end of time T. Time T may be, say, ten years,
and the half-life of the atom 1 nsec. Such a belief is of course unprovable
because, anyhow, somebody has to look at the recorder; and even if the
recorder includes a recording clock, it could be logically maintained that the
transition has been induced, via advanced waves, by the final look of the
observer.
Thus it seems that the very consistency of the London-Bauer philosophy,
which I have built into mine, implies that our world is full of rudimentary
psyches which (as proved by the preponderance of retarded over advanced
waves) are usually more passive than active, more of the sort of impartial
observers than of energetic wills. However, the truly wonderful facts of both
biological ontogenesis and phylogenesis may well suggest that at least some
among these rudimentary psyches are more willingly inclined. I certainly
52 O. COSTA DE BEAUREGARD

need not recall that quite a few very eminent biologists, philosophers, or
even mathematicians have made this sort of speculation; there are far more
names here than just the two I have quoted ([27], [28]).

9. CONCLUDING REMARKS

Concluding, I agree with those distinguished physicists and/or philosophers


of science who do not see, inside the quantal formalism, anything akin to
the stochastic event that was postulated to occur in classical statistical
mechanics. Therefore, with von Neumann and with London and Bauer,
I feel that a special postulate is necessary for bringing into existence the
stochastic event, or transition, which quantum mechanics certainly needs on
experimental grounds. And I cannot see any other plausible way of doing
this than by stating that the so-called col/apse of the state vector occurs
through an act of consciousness on the part of the observer.
Then, arguing from the philosophy of physical irreversibility that is
today accepted by many thinkers, and from the corollary I believe should be
drawn from intrinsic time symmetry to intrinsic symmetry between
cognizance and will, I am led to conceive that the act of consiousness pro-
ducing the quantal transition can be an act of will just as well as an act of
cognizance; that is, the sink of an advanced wave just as well as the source of
a retarded wave.
Finally, the need for consistency of the whole scheme leads me to think
of the world we are living in as a Leibnitzian world, where cats are rather
high in the hierarchy of monads.
Through the space-time vacuum quantal information waves ripple, with
full relativistic covariance, from monad to monad, and they are de jure just
as alive in their advanced interpretation as they are so obviously cascading
in their retarded interpretation. Paraphrasing Bergson, I would say that
advanced waves are rather dormant than absent.
Of course, I am well aware that, in proposing this high-brow sort of
paradox, by referring to Leibnitz and to an updated version of the Loschmidt
and Zermelo paradoxes, I am ipso facto letting in a paradox of the ~xtreme
low-brow, creeping sort: psychokinesis.
Let me summon Hippocrates as an attorney, because of his aphorism: 16
Extreme remedies are the most appropriate for extreme diseases. Fifty years
of writing (more than once by competent and/or subtle thinkers) without
having settled matters certainly proves that, notwithstanding its vigorous
health, quantum mechanics suffers from an enigmatic illness and needs an
TIME SYMMETRY AND QUANTUM MECHANICS 53

appropriate operation. If the reader thinks the treatment I am offering is


somewhat akin to acupuncture, please consider that Everett has seriously put
forth something even more fantastic, and a theory which (as far as I can see) is
not falsifiable. My theory, if admittedly less respectable when seen from
below than when seen from above, is at least falsifiable.

Institut Henri Poincare, Paris, France

NOTES

1 The argument has been rediscovered independently by Abele and Malvaux (15).
2 Yihnaz points out that the Galilean group formula does not preserve orthogonality of
light rays and wave planes in ordinary space. But this orthogonality is preserved by the
relativity of simultaneity. Incidentally, a very sinlilar argument answers Lande [56)
when he states that the Einstein-de Broglie formula p = I'Ik is not invariant under the
Galileo transformation.
3 Poincare is the proponent of the four-dimensional interpretation of relativity, and
Minkowski's inspirer.
4 That the quantal formalism has nothing in itself to tell us that an individual event (or
transition) occurs can be displayed in more than one way. Here is the simplest one.
Consider the expansion of the state vector upon the orthogonal set tPK characterizing a
measurement process t/J(x, t) = I: CK(C)tPK(x), where I cKi2 is the probability of finding
the state tPK. There is nothing inside the formalism implying that some sort of discon-
tinuity exists and induces the transition. Thus most authors oppose the continuous or
causal development of t/J, as governed by Schrodinger's equation, to the discontinuous
jump that the 'collapse of the state vector' must be postulated to be.
5 For an extensive bibliography see [23).
6 It does not seem plausible that macroscopic physical irreversibility has its root in the
rare and weak T-violating interactions that have been recently discovered. Moreover,
contrary to Lee and Yang's C-violations, the T-violations are not yet well understood.
It is possible that, after all, they fall in the general category of time asymmetry as
~overned by a boundary condition. -
To say that the A ~ B and the B ~ A transitions have the same predictive (intrinsic)
probabilities is not identical to saying that th.e (intrinsic) predictive probability that A
goes into B equals the (intrinsic) retrodictive probability that B has come from A. That
these two sorts of reversibility should be equal is known as the principle of detailed
balance. This principle holds in many cases, for instance, in the two that are quoted.
8 For an extensive bibliography see [23).
9 Cox [29) gives an extensive bibliography; see also the references in (21)- [23 J•
10 One need not say that entropy is an increasing function of probability if, and only if,
the basis of logarithms is larger than one.
11 In fact 'Boltzmann's constant' k was defined by Planck in the same historic paper
where he proposed his h constant. •
54 O. COSTA. DE BEAUREGARD

12 Such a distinction loses its objective testability if we are speaking of one individual,
quantal transition. Then it has solipsistic significance only. Let us display the (explicitly
covariant) mathematical formalism underlying the philosophical problem.
According to Dirac and to Lande the composition .law of quantal probability am-
plitudes may be written as

(a I b) (b I c) (c I a') =6 (a, b') (1)

where 6 denotes the Kronecker delta (6 = 1 if a =a'; 0 otherwise), the three expressions
(1) are probability amplitudes, and are such that

(2)

An appropriate summation or integration is implied by repeated symbols at each junc-


tion, such as I b) (b I. In fact, these junctions are projection operators.
Let us take an example: <x Ix') may be the propagator, or relativistic Green's func-
tion, associated with the wave equation under consideration, and <x I k) = (k I x)*
the function exp(ikx) on the mass shell, 0 otherwise; kx denotes the space-time scalar
product of the point instant x and the 4-frequency (or propagation vector}k. This
formula is 'manifestly covariant'. My book [36) displays quite a collection of covariant
formulaS of this sort, for reciprocal Fourier transforms, etc.
Formula (1) may of course be written as

(a I c)=(a I b)(b I c) (3)

and a combined use of formulas (1) and (3) yields

(a I b)(b la')=(a I c)(c la')=6 (a, a') (4)

that is, conservation of orthonormality as eXfressed by summation either over b or over c.


This is true in particular if b = x and c = x , that is, if (b Ic ) is the propagator. Con-
servation of orthonormality thus occurs modulo that the wave equation is obeyed, and
this is because, in this formalism, obedience to the wave equation is built into the very
defInition of all of the algorithms. For this I refer the reader to my book [36).
It shOUld be noted that this automatic conservation of the norm and the orthogona-
lity is due to the fact that the composition law (1) or (3) is for probability amplitudes
and not for probabilities. This remark has a strongly Landeian flavor, and points toward
very deep implications, in natural philoSophy, of the highly specific probabilistic for-
malism of the quantum theory (which of course is a leitmotiv in this paper). It should
also be obvious that the preceding formulas are completely symmetric with respect
to past and future, that is, prediction and retrodiction.
13 This point is also made by d'Espagnat ([7), p. 302).
14 See (46), (47) (especially pp. 196-197), and (48).
1 5 It thus seems that any operational hidden variable theory would be even more para-
doxical than the accepted quantal formalism. This certainly echoes another famous
contest between modelism and formalism, three quarters of a century ago.
16 Reece [55) gives this quotation (p. 88). This review article contains numerous
references.
TIME SYMMETRY AND QUANTUM MECHANICS 55

REFERENCES

[1] Jauch J. M., 1968. Foundations of Quantum Mechanics (Reading, Mass.: Addison-
Wesley).
[2] Krips, H. P., 1969. Philosophy of Science 36 145.
[3] Mo1dauer, P. A.,Phys. Rev. D 5 (1972),1028.
[4] Moldauer, P. A., Found Phys. 2 (1972),41.
[5] Ludwig, G., 'The Measurement Process and an Axiomatic Foundation of Quantum
Mechanics.' In Foundations of Quantum Mechanics, ed. by B. d'Espagnat, p. 287
(New York: Academic Press).
[6] Sharp, D., 1961. Philosophy of Science 28 225.
[7) d'Espagnat, B., 1971. Conceptual Foundations of Quantum Mechanics (Menlo
Park: Benjamin).
[8) d'Espagnat, B., 1971. 'mesure et non-separabilite.' In Foundations of Quantum
Mechanics, ed. by B. d'Espagnat, p. 84 (New York: Academic Press).
[9) Wigner, E. P.,1971. 'The Subject of Our Discussions.' In Foundations of Quantum
Mechanics, ed. by B. d'Espagnat, p. 14 (New York: Academic Press).
[10) Hooker, C. A., 1970.Am. J. Phys. 38851.
[11) Hooker, C. A., 1971. Philosophy of Science 38224.
[12) Hooker, C. A., 1971. Philosophy of Science 38418.
[13 ) Hooker, C. A., 1973. 'The Nature of Quantum Mechanical Reality: Einstein versus
Bohr.' In Paradox and Paradigm, ed. by R. G. Co1odny (Pittsburgh: Univ. of
Pittsburgh Press).
[14) Hadamard, J., 1930. Cours d'Analyse Professe a l'Ecole Poly technique, Vol. 2,
p. 385 (Hermann).
[15) Abele J. and P. Malvaux, 1954. Vitesse et Univers Relativiste (paris: Sedes).
[16) Yilmaz, H., 'Could Relativity Theory Have Been Discovered in 1728?' (mimeo-
graphed paper).
[17) Poincare, H., 1906. 'Sur la dynamique de i'electron.' In Rendiconti Circolo
Matematico di Palermo 21 129.
[18) Colodny, R. G. (ed.) , 1973. Paradox and Paradigm (Pittsburgh: University of
Pittsburgh Press).
(19) von Neumann, J., 1932. Mathematische Grundlagen der Quantenmechanik
(Berlin: Springer).
(20) London F. and E. Bauer, 1939. La Theorie·de l'Observation en Mecanique Quan-
tique (paris: Hermann).
[21) Costa de Beauregard, 0., 1964. 'Irreversibility Problems.' In Proceedings of the
1964 International Congress for Logic, Methodology and Philosophy of Science,
ed. by Y. Bar-Hillel, p. 313 (Amsterdam and New York: North-Holland).
(22) Costa de Beauregard, 0., 1970. 'Is There a Paradox in the Theory of Time Asym-
metrY.' In A Critical Review of Thermodynamics, ed. by E. B. Stuart, B. Gal-Or,
and A. J. Brainard, p. 461 (Baltimore: Mono Books Corp.).
(23) Costa de Beauregard, 0., 1971. Studium Generale 24 10.
(24) Lande, A., 1965. New Foundations of Quantum Mechanics (Cambridge: Cam-
bridge University Press).
(25) van der Waals, J. D., 1911. Phys. Z. 12547.
(26) Gibbs, J. W., 1914. Elementary Principles in Statistical Mechanics, p. 150 (New
Haven: Yale University Press).
56 O. COSTA DE BEAUREGARD

(27) Bergson, H., 1907. L 'Evolution Creatrice (Paris: Presses Universitaires de France).
(28) Fantappie, L., 1944. Teoria Unitaria del Mondo Fisico e Biologico (Rome:
Humanitas Nova).
(29) Cox, R. T., 1961. The Algebra of Probable Inference (Baltimore: The Johns
Hopkins Press).
(30) Brillouin, L.,1962. Science and Information Theory, p.294 (New York: Academic
Press).
(31) Fock, V., 1948. DAN SSSR 60 1157.
[32] Ritz W. and A. Einstein, 1909.Phys. Z. 10323.
[33] Jaynes, E. T., 1957. Phys. Rev. 106620; 108 171.
[34] Elsasser, W. M.,1937. Phys. Rev. 52987.
[35] Riesz, M., 1946. 'Sur Certaines Notions Fondamentales en Theorie Quantique
Relativiste.' In Actes du Dixit:me Congres des Mathematiciens Scandinaves
(Copenhagen: Gjellc;rup).
[36] Costa de Beauregard, 0., 1967. Precis de Mecanique Quantique Relativiste (Paris:
Dunod).
[37] Wightman A. S. and S. S. Schweber, 1955. Phys. Rev. 98812.
[38] Dirac, P.A.M.,1930,1935, 1947. The Principles of Quantum Mechanics (Oxford:
Clarendon Press).
[39] Schmidt, H., 1970. J. Appl. Phys. 41462.
[40] Schmidt, H., 1970. J. Parapsychol. 34175.
[41] Price, G. R. 1972. Science 175 359.
[42] Einstein, A., B. Podolsky, and N. Rosen, 1935.Phys. Rev. 47 777.
[43] Einstein, A., 1928. 'Electrons et Photons.' In Rapports et discussions du 5e Con·
seil Solvay, p. 253 (Paris: Gauthier Villars).
[44] SchrOdinger, E., 1935. Naturwiss. 33807,823,844.
(45) Renninger, M.,1963.Phys. Z. 136251.
[46] Costa de Beauregard, 0.,1965. Dialectica 19280.
(47) Costa de Beauregard, 0.,1960. Dialectica 22187.
(48) Costa de Beauregard j 0., 1970. 'Discussion on Temporal Asymmetry in Thermo-
dynamics and Cosmology.' In Proceedings of the International Conference on
Thermodynamics, ed. by P. T. Landsberg, p. 539 (London: Butterworths).
(49) Bohm, D., 1951. Quantum Theory, Ch. 22 (New York: Prentice Hall).
[SO) Bell, J. S.,1965.Physics 1195.
[51) Shimony, A., 1971. 'Experimental Test of Local Hidden-Variable Theories.' In
Foundations of Quantum Mechanics, ed. by B. d'Espagnat, p. 182 (New York:
Academic Press).
[52) Freedman, S. J. and J. Clauser, 1972. Phys. Rev. Lett. 28938.
[53) Freedman, S. J., 1972. 'Experimental Test of Local Hidden Variable Theories,'
Ph.D. Thesis, unpublished, and references therein.
[54) Prosperi, G. M., 1971. 'Macroscopic Physics and the Problem of Measurement in
Quantum Mechanics.' In Foundations of Physics, ed. by B. d'Espagnat, p. 97
(New York: Academic Press).
[55) Reece, G., 1973. Int. J. Theor. Phys. 781.
[56] Lande, A., 1973. Quantum Mechanics in a New Key, p. 11 (New York: Exposi-
tion Press).
HANS EKSTEIN

IS PHYSICAL SPACE UNIQUE OR OPTIONAL?

1. INTRODUCTION AND SUMMARY

Ever since H. Poincare [1] asserted the conventional nature of the geometry
of physical space, there has been a debate on this problem, particularly since
a new kind of space was introduced by Einstein's General Relativity.
Accounts and bibliographies were given by many authors, among which we
cite A. Griinbaum [2] and J. Ehlers [3].
There are two concepts of physical space-time. One, SF' is that of a ftxed
arena in which events take place. The other, SD' is that of a space-time
shaped by events. The second depends on the state (initial conditions) or on
the external fteld, the ftrst does not.
The main assertions of the present paper are:
(1) The ftxed space-time SF is neither incompatible with, nor made
superfluous, by Einstein's theory. SF is experimentally explorable, unique,
and probably identical to Minkowski space M.
(2) The dynamical space SD is largely optional. It can be chosen to be M,
but the natural choice is Einstein's pseudo-Riemannian manifold.
The claim that the General Theory of Relativity (GTR) has not made
the concept of the fixed space SF untenable is based on a paper by the author
in collaboration with Y. Avishai [4]. This paper, starting from a ftxed flat
space, arrives at results eqUivalent to the general theory of relativity, not only
(as was done previOusly) with respect to speciftc gravitational effects ([5],
[6]), but it derives from special relativity and the acceleration group an
operationally testable version of Einstein's equivalence principle.
Poincare [1] has analyzed qualitative sensory perceptions to derive some
of the properties of physical space (SF)' namely those summarily described
by the words homogeneity and isotropy. As an extension of this approach,
the present paper describes a quantitative method by which the automorphisms
of the algebra 0 of observation procedures are experimentally explored.
Those transformations of SF that induce automorphisms of 0 are, by a
deftnition substantially due to Poincare, the automorphisms of SF. Evidence
favors the Poincare group and Minkowski space M as the choice for SF.
This choice seems to meet the following objection.

57
R. S. CohenandM. W. Warto/sky (eds.j, Physical Sciences and History o/Physics, 57-74.
© 1984 by D. Reidel Publishing Company.
58 HANS EKSTEIN

The existence of an invariant pseudo-distance! in Minkowski space can be


used to test the choice of M for the space SF" Indeed, if the assertion SF = M
is true, there must exist observation procedures whose outcomes have
(approximately) the relations attributed to mathematical invariant pseudo-
distances.
It would be tempting to adopt, as a standard of time-like distance measure-
ments, atomic clocks as in non-gravitational physics (Le., corrected for
electric or magnetic, etc., but not for gravitational effects). However, the fact
of gravitational red-shift (predicted by Einstein) shows that this choice is
inconsistent.
If an atomic emission frequency is assumed to provide a valid time
measuring device even in a gravitational field, a purely empirical argument by
Schild [7] shows that time-interval measurements are not translation-
invariant. However, an atomic clock corrected by a factor that depends on
the gravitational potential can serve as a translation-invariant timepiece.
Needless to say, any actual measurement requires a host of precautions and
corrections which challenge the experimental physicists. However, his success
is definitely testable.
If this kind of observation procedure for a translation-invariant time
interval is adopted, the red-shift is interpreted as showing a change in the
frequency of a given atomic emission line caused by a translation in a gravita-
tional field. Hence, the objection against SF =M can be put aside, at the cost
of some complications.
There is a long-standing controversy about space-time distances in the
literature. The positivist's argument runs roughly as follows: Einstein cut the
Gordian knot of speculations on the nature of time with the simple assertion:
time is what a good clock measures. This positivist attitude, which was instru-
mental in building the Special Theory, was also the basis of the General
Theory. There again, distances are defined as that which is measured by
(properly adjusted) clocks, light signals, yard-sticks. The weakness of this
definition was pOinted out by Einstein himself [8], in his discussion with
Reichenbach: the apparently harmless qualifications "good" or "properly
adjusted" hide the fact that the concept of distance cannot be isolated from
the context of all other physical processes and concepts.
To sum up the discussion on this point: the concept of distance used in
this paper is a kind of afterthought rather than a primary operational object,
and the instruments and associated software used to measure it are not given
at the outset but must be found either by theory or by trial and error. The
non-tautological assertion is that they can be found.
IS PHYSICAL SPACE UNIQUE OR OPTIONAL? S9

To introduce the dynamical space SD, we consider first special relativity


without gravitation. The physical predictions of a theory are contained in the
morphism from the algebra .0 of observation procedures onto the algebra III of
observables. For non-gravitational classical fields, this algebra \!l is frequently
taken to be the algebra of functionals on a class of functions (the solutions of
the equations of motion) on Minkowski Space M = SD. However, other
faithful representations of the same algebra are possible, without any change
in the observable predictions of the theory. In particular, the subcarrier space
M may be replaced by any space ~ that has the cardinality ofM. A sub-carrier
space ~ has some similarity with SF and can be considered as a possible
dynamical physical space SD if it is at least locally homeomorphic to SF.
In the presence of gravitational fields, Einstein's pseudo-Riemannian mani-
fold has proved to be enormously successful. It is the natural choice for SD,
just as M is natural for non-gravitational physics.
Poincare says that asking: can physics be eqUivalently described in
different spaces? is similar to the question: can geometric relations be
expressed in feet as well as in centimeters? Indeed, in our view, the different
possible spaces SD characterize isomorphic representations of the same
algebra. Their 'conventional' or rather optional nature is evident.
In particular, the space SD' on which equations of motion are defmed and
which varies from experiment to experiment, can be chosen so that its metric
tensor has always the usual relation to measurements with standards not
corrected for gravitational effects. The great power of a theory formulated in
this space is obvious.
Is there, then, a reason for using the fixed space SF? An explicit and
general physical interpretation of Einstein's theory seems to require this
space. The usual references to moving observers seem obscure to many of us.
The paper mentioned as reference [4] uses flat space to make precise and
general what was allusive and illustrative.
The scope of the present study lies in reproducible, controlled laboratory
experiments. This explains, at least in part, why the conclusions here differ
from those of most authors on General Relativity. They have in mind situa-
tions with far stronger limitations on the explorability of nature. The spirit of
this paper is close to that of S. Weinberg's book [12], especially pages 147-
148.

2. GENERAL ASSUMPTIONS

A controlled experiment requires shielding or isolation of the laboratory from


60 HANS EKSTEIN

uncontrolled external influences. 'This control does not exclude external


fields, magnetic, gravitational or others, if they are controlled and reproduced
in each repetitive experiment. The test for successful control is causal
behavior within the laboratory. If initial conditions within the laboratory
determine the future (at least statistically) we infer that external objects have
a controlled and reproduced influence on the laboratory objects.
An experiment has three stages: (l) a state-preparing procedure which
terminates at an 'initial' time when the produced system ceases to interact
with the preparing instruments; (2) a time during which the produced system
evolves without interacting with preparing or observation instruments; and
(3) an act of observation which results in the printout of a real number.
s
State-preparing procedures are not, in general, pure. An accelerator that
emits altematingly mesons and protons in a random manner, is an example
of a state-preparing procedure.
Procedures can be considered as pairs (b, {Pn }) consisting of blueprints
b for the construction of apparatus, and a set {Pn } of events (points). For
instance, b is the blueprint of a telescope and {Pn } are points on world lines
of marks on the telescope. With a slight abuse of language, we identify the
physical act of performing a procedure with the instruction or computer
program for its performance.
Observation procedures are, in general, destructive, i.e., the evolution of a
state is violently altered by the act of observation. For instance, verifying that
three bodies are on a straight line can be done by light rays if there are no
obstructing or perturbing bodies or fields between them. Otherwise, the
experimenter may remove the sources of fields and use light rays, possibly
correcting for the displacement of the three tested bodies by the act of
removal of the fields.
An observation procedure ex uses several instruments alternatively, e.g.
different types of scales for different orders of magnitude of masses to be
weighed. The instruction will call for the use of indirect methods, depending
on the state. To measure the angular distance between two ~stant objects,
one can use a telescope if the light path is unobstructed. If it is obstructed by
fields, one can: (l) correct for their effect if the fields are known, or (2)
repeat the measurement when the field is absent (as after a solar eclipse)
provided one has reason to believe that the angular distance has not changed,
or (3) remove the source of the disturbing field quickly by the use of a high-
energy projectile. (Extrapolating the increase of energy of accelerators, we
expect this latter method to become available shortly!)
All such possiblities are supposed to be contained in the instructions
IS PHYSICAL SPACE UNIQUE OR OPTIONAL? 61

accompanying blueprints (b). The mean value of many (ideally infinitely


many) repeated experiments with a given state-preparing and observation
procedure results in a real number, the expectation value. The set of these
expectation values depends on the external field and the nature of the system;
in theoretical terms, on the Hamiltonian or the equations of motion. Let§'
be the set of state-preparing procedures s, and 0 the set of observation pro-
cedures ii. An expectation '€ is the map

'€:YXD-+R
from the cartesian product of Yand D into the reals such that '€ (S, ii) is the
~

mean value of observations ii on a given system with state-preparing proce-


dure s.
Procedures ii can be divided into equivalence classes a such that those
procedures which give equal expectation values for all expectations € and
all state-preparing procedures s are equivalent (-). We call the quotient
set .0 /_ = .0 the algebra of (equivalence classes ot) observation procedures.
The algebraic structure will be introduced presently.
Similarly, those state-preparing procedures s which give rise to equal
expectation values € (S, ii) for all observation procedures ii and all expecta-
tions € are eqUivalent (-). We define the quotient set by §'/_ = Y, and
call it (pending justification), the convex linear set of (equivalence classes
ot) state-preparing procedures s.
Given a program for the performance of a procedure, a new program can
be written by the following alteration: given a transformation g of space-time
SF, such that each point P is carried into gP by

g: P l-+gP, ,

one associates a new program (b, {g Pn }), which each procedure ii:
(b, {Pn }) 1-+ (b, {g Pn}). (2.1)
In general, the new program will not be implementable physically. The
implementable transformations g will be called motions, and they induce
transformations

and

(2.2)
62 HANS EKSTEIN

if equivalence classes of procedures are preserved by the map (2.1). Thus far,
the framework is equally valid for quantum mechanics and classical theory.
Oassical physics assumes that, in the usual language, two observations can
be 'performed simultaneously' without disturbing each other. More precisely,
a single sample (the product of a single act of state preparation) can be
subjected to two observation procedures without interference. The opera-
tional meaning of this expression is the equality of mean values of observa-
tion results and of statistical distributions of outcomes (1) in a sequence of
experiments where only one procedure is used on one sample and (2) in a
sequence with two observation acts used on a single sample, the state-
preparing procedure and the external field being the same in the two
sequences. The most frequently used example refers to two simultaneous
observation acts such as position and momentum measurements, but an
equally good example is that of two successive position measurements on the
same sample. From here on, we consider only classical physics.
Of course, this non-interfering nature of two observations is asserted only
for some (very gentle) observation procedures, such as position measurements
with very soft X-rays, but the assumption alleges that such gentle procedures
exist in each equivalence class of observation procedures for all pairs (Ii, ~)
s
and for all state-preparing procedures and all expectations e. One can then
consider the mean value of the sums of individual outcomes {s'n (an,
{s'n (~)},
1 N I I -
lim N ~ [s n (ii) + Sn (ft»), (2.3)
N~" 1
obtained in such experiments involving two measurements on each sample
and assert that it is the sum of the two mean values
1 N 1 N -
lim N ~ sn (Q) + lim N ~ Sn (ft), (2.4)
N-+oo 1 N~oo 1
obtained in the two sequences involving only one observation (a or ~) on
each sample.
This addition of individual outcomes defines operationally a procedure
(a, 'jf)+ such that
(2.5)

This property of the (not ordered) pair (a, if)+ suggests that addition be
defined in o. By construction, the mathematical image of the procedure
IS PHYSICAL SPACE UNIQUE OR OPTIONAL? 63

(a, ~)+ is the element a + (3 of a linear set O. With the definition 'f(s, a) =
€ (s, a) and with the convention that € is linear in its second argument,
Equation (2.5) now reads

€ (s, a) + € (s, (3) =€ [s, (a + (3)]. (2.6)

The same considerations apply to the multiplication of the terms of two


sequences of outcomes s'n (a) and s'n (~) of observation procedures a and
(3, if observations are made on the same sample. The products ~'n (a) . s'n (~)
are defmed to be outcomes of a new procedure designated by the (not
ordered) pair (a, ~)A' The mathematical counterpart of this procedure is the
commutative product

0t{3 = (3a. (2.7)

Thus, the set 0 has the structure of a real Abelian algebra, and each algebraic
operation has a well-defmed operational counterpart in .0 .
State-preparing procedures can bctcomposed in a well-known marmer. (See
[10].) Thereby, the quotient set Y /'" acquires the structure of a convex
linear set.
We have considered two types of alterations of procedures: the replace-
ment of points P by their images gP such that equivalence classes in .0 are
preserved, and the addition and multiplication of individual outcomes of two
procedures performed on the same sample. These two types of alterations
commute; they can be performed in reverse order without changing the
result. Hence, a motion g that preserves equivalence classes in .6 induces an
automorphism in 0 :

Vg :0 ~O·

3. AUTOMORPHISMS OF THE PHYSICAL SPACE SF

Most inquiries into the nature of physical space-time concentrate on its


metric or congruence properties, possibly infmitesirnal. Our approach is in the
line of tradition begun by Klein's Erlanger Program, and continued by S.
Lie [9] and Poincare [1] in which the group of automorphisms defines the
space and its binary invariants. From this viewpoint, the metric or pseudo-
metric of SF carmot be freely defmed by an operational rule - it is not
64 HANS EKSTEIN

optional or conventional. While the knowledge of the group does not give an
instruction for the measurement of an invariant distance, the invariance of
the outcome provides a conclusive test for the validity of a distance mea-
suring procedure.
An operational definition of physical space-time means a set of rules by
which outcomes of measurements are related to statements about the
structure of a mathematical space SF. As an example of such an enterprise,
we cite Poincare's rule [1] for the exploration of space structure by sense-
perception and bodily motions.
Some phySicists do not think that such an attempt is useful or can be
successful. Einstein; in his later years, thought that only the success of a
whole theory, including geometric as well as dynamical assumptions, can be
considered as corroborative evidence for the assumed geometry. Disregarding
such warnin~, I will attempt to explore SF without dynamical assumptions
with a view to establishing principles into which many different dynamical
theories can be fitted - as special relativity allows both the Maxwell and the
Born-Infeld electromagnetic theories to be fitted.
Before proposing a formal operational definition of SF, we consider in
informal language some physical facts that suggest properties of SF. We have
already impliCitly made one assumption of this kind: that a sufficiently small
region R in space-time (or a limit of a sequence of such reiions) can be
uniquely identified with a mathematical point P so that a statement such as:
the electromagnetic field vanishes at P, is meaningful and, in principle, veri-
fiable. I know that some authors question this assumption, especially if P
is in an empty space-region, but it is so widely accepted that no further
justification will be given.
The possibility of repeating the same measurement at different times and
places requires that different points P lo P2 should have intrinsically equal
properties, i.e., that physical space-time should be homogeneous. In
mathematical language, homogeneity of a space means the existence of a
transitive group of automorphisms, i.e., every point PI can be mapped onto
any point P2 by some element of the group. We want to point out that the
intuitive notions of sameness and homogeneity correspond precisely to the
mathematical definition if a link between automorphisms of SF and the
class-preserving motions of observation instruments is postulated.
If SF is homogeneous, one expects that for any pair PI, P2 of points, there
exists an automorphism g such that
IS PHYSICAL SPACE UNIQUE OR OPTIONAL? 65

and that the transformation

(b, {Pn }) 1-+ (b, {g Pn })

preserves equivalence classes of procedures in .0 and therefore induces an


algebraic automorphism in D.
The next remark, based on elementary experience, is that motions gl and
g2 can be performed sequentially to result in a new motion gh2, and that the
motion g can be reversed to give a new motion g-l with the property g-l g
=gg-I =I. Thus, the class-preserving motions form a group G.
Needless to say, there is overwhelming evidence in favor of operational
tests of homogeneity. What is more, the positive outcome of these tests is
really a necessary condition for the possibility of physical research. If it were
not possible to have a set of (e.g.) temperature-reading instruments which give
concordant reading; after transport or time-lapse, how could one rely on a
measurement at all?
There is an apparent contradiction between the homogeneity postulate
and the GTR. For instance, the sum of angles in a triangle varies as the obser-
vation procedure is moved through space - algebraic relations between pro-
cedures are apparently not preserved by motions. However, the observation
procedures that correspond to the metric-induced observables of the GTR
using natural clocks, light signals, and rods, uncorrected for gravitational
effects - do not have simple geometric relations. More explicitly, the angles
a; between light beams forming the sides of a triangle satisfy the relation
3
~ e(s, 0:;) = 1T (3.1)
;=1
for some states and expectations e, but not in the presence of a gravitational
field. One cannot infer the relation

(3.2)

as an equation between procedures o:;E.o. It is not surprising that Equation


(3.1) is not invariant under any motion - i.e., that there exists in general no
motion g of SF such that the transformation

Vg : (b, {Pn }) 1-+ (b, {g Pn }) (3.3)

induces the relation

~ e (s, Vg 0:;) = 1T (3.4)


66 HANS EKSTEIN

In our terminology, the procedures Otj would not be acceptable as valid


members of the set.o of observation procedures. On the other hand, proce-
dures {jj that use gravitational corrections on light beams to determine the
(straight) sides of a triangle, satisfy Equation (3.1) for all states and external
fields so that
!; {jj = 'IT

is legitimate and the analog of Equation (3.4) holds.


Continuing our preliminary inquiry into the structure of SF, we consider
more closely the restrictions imposed on it by the postulate that measure-
ments be repeatable. The intuitive concept of repeating an observation act at
PI at a later or spatially displaced point P2 will now be made more precise.
There are many automorphisms g E G that carry the space-time position of a
mark PI on the observation instrument onto P2 , but only one of them will
lead to the same measurement at P2 - e.g. the measurement of the z-com-
ponent of a vector field at P l will be repeated atP2 only if the instrument is
translated and not rotated or boosted.
Hence, we consider a minimal transitive subgroup T C G such that there
exists a unique element t E T that produces the repetition of the measure-
ment for given points PI andP2 ESF.
I.et
ex = (b, {Pn })

be an observation procedure, gET an automorphism of SF, and Po and


g Po reference points of the two measurements: e.g. the world points of
the left upper comer of an instrument box described by b. Consider a point
Pm E {Pn }. Since the group T of automorphisms is transitive, there exists a
unique automorphism hE Tsuch that Pm =hPo . If the procedure (b, {gPn }}
is to be a faithful repetition of (b, {Pn }}, the point g Pm must be - in loose
language - the same with respect to g Po as Pm was with respect to the
reference point Po. In other words, g (h Po) = h (g Po), i.e., the automor-
phisms g and h commute. (See Figure 1.)
Experimental physicists spend much of their time checking their equip-
ment after a lapse of time or transportation. They verify that different ways
to observe the same quantity still give identical results. In our language, they
verify the preservation of equivalence classes in .0 under certain time or
space displacements. They also check the preservation of algebraic relations
between instrument readings, e.g. a galvanomenter (J) and a Joule's heat
measuring instrument (J2 R).
IS PHYSICAL SPACE UNIQUE OR OPTIONAL? 67
h9Po

Po
Fig. 1. In the usual elliptic language, the point hgPo characterizing the second measure-
ment is "the same with respect to gPo" as the point hPo characterizing the flIst measure-
ment is "with respect to Po".
In our language, they verify that certain space or time displacements are
motions g that preserve algebraic relations between classes of observation
procedures and hence induce automorphisms of the algebra O.
These heuristic arguments suggest the defmition: those transformations g
of SF that induce class-preserving transformations of the set .6 of observa-
tion procedures and hence induce automorphisms of the algebra 0, are auto-
morphisms of SF·
The space SF itself is the space of cosets (with respect to stabilizers at a
point) of this group.
A similar definition was used by Poincare [1] to determine the nature
of physical space on the basis of sense perceptions and bodily displacements.
Let us recall that the equivalence between procedures refers only to universal
agreement of expectation values, not to those that depend on the state or the
external field. As an example, two procedures, one measuring the energy, the
other mass X (velocity)2/2 agree with respect to those states where a particle
is free, but not in general. These two procedures do not belong to the same
equivalence class a E O.
Determining motions by observation is not a trivial task. A test for the
acceptability of an observation procedure is its membership in an equivalence
class of procedures a E 0 that is preserved by a motion (e.g. two thermo-
meters made according to an identical program and sufficiently close must
give identical readings also after a lapse of time or after a displacement). The
operational determination of automorphisms of SF is, in a sense, a circular
process because the validity of a procedure requires transformation properties
induced by automorphic transformations of space-time. The non-tautological
68 HANS EKSTEIN

statement is, of course, that such observation procedures and transformations


of SF can be found within a good approximation. The operational test of an
assertion about automorphisms is thus a matter oftrial and error.
It is all the more important that there exists a general argument to show the
necessity of the existence of a transitive subgroup of commuting automor-
phisms of SF. This argument was based on the very general postulate that
manipulative physics, with repeatable measurement procedures, exists. I
would like to make a remark about the philosophical nature of this argument.
It claims that the possibility of acquiring knowledge about nature implies a
restriction on the structure of space-time. It is similar to Kant's argument
that our ability to conceive any but Euclidean space implies that the intelligi-
ble world must ha~e this structure. My argument differs in that it is not based
on a claim of the non-existence of mathematical concepts, and also in that
it does not imply by itself the flatness of SF' as will be shown presently. The
properties of SF that follow from this neoKantian argument are:
(1) There exists a subgroup T C G of automorphisms of SF that acts
transitively.
(2) Given two points PI, P 2 € SF, there exists a unique t € T such that tP 1
=P2 •
(3) T is Abelian. Oearly, these properties are insufficient to derme SF
uniquely. I.et us consider what additional information can be obtained from
qualitative observations. We infer from general experience:
(4) There exist 4 one-parameter subgroups T; of T (i.e., subgroups such
that an element t € T is a power tg of an element to € Ti) that generate T-
i.e., every element t € T equals a product ii 1 ti with ti € Ti.
(5) t € T can be generated by squaring, i.e., there exists an element r € T
such that t =r2 •
These 5 conditions sugg~t the inference that SF is a 4-dimensional mani-
fold.
That the distinguished Abelian group T of automorphisms considered
above is a translation group, is a less certain assumption, but hardly contro-
versial, at least as an excellent approximation.
Finally, the existence - at least locally - of a metric or pseudometric
invariant under automorphisms, is supported by overwhelming experience.
We can then conclude, from the elementary theory of manifolds that SF is
flat, and can be taken to be a 4-dimensional affme space V4 •
While these conclusions were obtained on the basis of a postulate indis-
pensable for the existence of physics and of general (nonspecific) observa-
tional facts, the determination of the full group of automorphisms is the
achievement of very detailed theoretical considerations by Einstein, and their
IS PHYSICAL SPACE UNIQUE OR OPTIONAL? 69

experimental verification. It is, to the best of our knowledge, the connected


Poincare group.
The existence of a bilinear invariant, its uniqueness up to a constant, and
hence of the Minkowski pseudo-metric follows then from the properties of
the group P of automorphisms of SF. In the presence of a gravitational field,
no simple operational counterpart of this pseudometric can be given. How-
ever, the invariance of the procedure under automorphisms of SF is a test for
its validity. Invariance of the procedure 6 (P lo P,.) means that in a state s
where there are distinguishable events at the world points Plo P2 , gPl and
gP2 (e.g. flashes of different colors),
€ [s, 6 (P lo P2)] =€ [s, 6' (gP1o gP2)]
where g is any automorphism of SF.

4. THE ALGEBRA'll OF OBSERVABLES

The concept of the algebra of observables is more familiar in the quantum the-
ory of relativistic fields than in classical theory, but it proves useful here too.
In the simple case of a classical nonrelativistic particle, the algebra consists
of all continuous real-valued functions F of the position x and the momentum
p. Addition and multiplication of these functions are defmed in the natural
manner.
The physical interpretation of these observables is given by the procedures
that are used to measure them, i.e., by their inverse image in the algebra .0 of
observable procedures. The morphism ell: .0 -+ \]( that carries procedures into
observables is many-to-one. For instance, the usual shorthand statement
"energy equals mv 2 /2" means, explicitly, that a procedure e (e.g. using a
calorimeter) gives, for all states of a free particle, the same reading as a pro-
cedure that measures the velocity, squares it and multiplies by m/2. In our
language,
eIle=m/2(cIlv)2 (e,vE.o) (4.1)
is the expression of what one can call a natural law.
This more explicit way to express the physical interpretation of a theory
is, of course, only pedantic and redundant in the simple case, but it seems to
be useful if not indispensable for theories such as Quantum Mechanics and
General Relativity.
All observable predictions of a theory for a given system in a given external
field are contained in the algebra m of observables and the (homomorphic)
map ell: .0 -+ 'll (the physical interpretation).
70 HANS EKSTEIN

The dynamics of a theory is contained in the automorphism QT: ~l -+- ~l


that is induced by a time-translation VT in .0 by the equation
(4.2)
For the simple case of a classical particle, 4>a is a continuous function FVc, p)
of position and momentum, and
QrFVc, p) =F[X (r),P(r)] , (4.3)
where X(r) and per) are the values of the functions X(t), andP(t) obtained
by solving the equations of motion with X(O) = x, P(O) =P as initial values.
This connection is illustrated by Figure 2. Among the isometries of Newtonian
space-time, we consider the time-translations {Tr }. They act on the algebra
of observables in two equivalent ways:
Aut (~) Aut (et)

150m (N) Hom (H)


Fig. 2. A time-translation isometry (x, t) -> (X. t.:t r) in N induces the transformation
Q = (b. {xn. t n }) 1-+ (b, {xn. tn + T}) of the set .0 of observatipn procedures, and this
induces an automorphism O! 1-+ Vr O! of the algebra .0 of (equivalence classes of) observa-
tion procedures. The compatible morphism <1>: .0 ..... ~l onto the algebra ~l of observ-
abIes induces the automorphism QT: ~l ..... ~l of ~l. On the other hand, the same iso-
metry of N induces a homeomorphism x 1-+ x (t), P 1-+ P (t) of the phase space II through
the equations of motion with the Hamntonian H. This homeomorphism of the carrier
space II induces, in turn, an automorphism of the algebra ~l of observables. This com-
mutative diagram is what makes the standard representation of ~l as the algebra of
continuous functions on II so convenient. In other possible representations of ~( this
convenience is lost.
IS PHYSICAL SPACE UNIQUE OR OPTIONAL? 71

(1) They induce automorphisms { Vt } in the algebra .0 of observation


procedures, which, in turn, induce automorphisms { Qr} of the algebra'll
of observables.
(2) They induce homeomorphisms x ~ X(t), p 1-+ pet) of phase space
II by the use of the equations of motion. These, in tum, induce automor-
phisms of the algebra of functions F through the transformation

F[x, p] 1-+ F[X (7),P(7)].

(2) is a realization of the abstract transformation (1).


Again this appears to be an unnecessarily complicated expression for the
simple rule: to calculate a function F of (x, p) (an observable) at time t,
calculate x(t) and pet) by Hamilton's equations with p(O) = p, x(O) =x, and
substitute these values into the function F.
Instead of looking at the algebra ~l as an algebra of functions F, one can
consider ~ as an abstract algebra, i.e., its table of additions and multipli-
cations. No observable information is lost by forgetting the concrete nature
of the algebra. Such an abstract Abelian algebra has infinitely many isomor-
phic representations by algebras of functions on carrier spaces n. The best
known non-standard representation, again for the classical particle, is the
space of solutions {X(t)} of Newton's equations. There is a one-to-one
map from phase space points (x, p) to world lines X(x, p) (t) with initial
values X(x, p)(D) =x,p(x, p)(O) =p. The algebra \'II is then represented as an
algebra of functionals on the (non-linear) carrier space W of world lines in
the Newtonian space-time N.
The realization of a time-transformation is now slightly different. A time-
isometry Tr of Newtonian space-time induces a homeomorphism

of the space W of world lines, which, in turn, induces the automorphism


~ -+ Q;
~ of the algebra of functions G (W) through

G[X(x,p)] 1-+ G[X(X(x,p)(7),P(x,p)(7))].

A third choice of n, useful as an introduction to General Realtivity [11],


is obtained by subsitituting a pseudo-Riemannian manifold SD as sub-
carrier space for the Newtonian space-time on which the world lines are
defined. This can be chosen so that the world lines on SD are geodesics, if
the force is purely gravitational. Again, this change of representation entails
72 HANSEKSTEIN

no observable consequences, but the equations of motion that select the


world lines are more suggestive.
In these non-standard representations of the algebra III , the simplicity
inherent in the commutative diagram (Figure 2) is lost, but something is
gained: in the last example, the simple statement that under purely gravita-
tional forces, particles move on geodesics.
For classical relativistic field theory, the generalization of the second
scheme described above is simple. Instead of world lines, we have fields, i.e.,
functions 1/1 on Minkowski spaceM, as elements of the carrier space, and the
algebra ~l consists of functionals on this carrier space n. As in the case of
world lines, the J:1elds are solutions of equations of motion. As in the simple
particle case, a time-translation induces a homeomorphism of n as follows:
Let the functions 1/1 be indexed by functions f on Euclidean 3-space such
that

I/Ij{x, 0) = 1(x), xEE,

i.e., fis the initial value of I/If. Then, the homeomorphism of n induced by
TT is the map

where

fT (x) = I/Ij{x, r).

For gravitation, the important step is a change of the carrier space n by sub-
stituting a differentiable manifold SD for the sub-carrier space M. The
differentiable manifold is then given a pseudo-Riemannian structure which
depends on the state considered. A short account of this procedure will be
given now.
The possibility of embodying Einstein's GTR is the present framework
is based on the following result of reference [4]: Given a field equation of a
vector field I/Iv on M, the effect of an added gravitational field is to add a
term of the form I"" VA I/Iv to each derivative a1/I/.I./axA , and similarly for
tensor fields. One can now map the sub-carrier space M onto a differentiable
manifold SD with the affine connection given by the symbols I"" VA. Locally,
of course, SD is homeomorphic to M, and smooth functions 1/1 on Mare
mapped onto locally smooth functions I{J on the differentiable manifold.
IS PHYSICAL SPACE UNIQUE OR OPTIONAL? 73

Globally, the nature of the occurrence of singularities on the functions f{J


cannot be determined in general; it depends both on the equations of motion
and the state (the solution). In the simplest case, the gravitational field is only
external, and the manifold is homeomorphic to M. Then, the space SD is
fixed, i.e., the same for all states, and smooth solutions l/J of M map on to
smooth functions of the differentiable manifold.
In general, the functions r JJ VA on M are determined by equations of
motion, and the pseudo-Riemannian space SD depends therefore on the
solutions l/J of the equations. The smoothness of the corresponding solution
on the differentiable manifold is obvious if SD is homeomorphic to M. In
general, the problem of singularities raises its ugly head. It has not been
explored in reference [4], and it may have to await more work on singu-
larities in the GTR.
The great advantage of Einstein's choice is the equality between the length
of time-like geodesics in SD and the length of time measured by an atomic
clock (uncorrected for gravitational effects) moving along that straight line in
M which is mapped onto the geodesic in question. While the usefulness of
Einstein's representation is manifest, it has no claim to uniqueness. One
could equally choose a one-dimensional sub-carrier space since only the
cardinality of two subrepresentation spaces has to be equal in order to derme
a one-to-one map of functions on them. Of course, the functions would then
not even be piecewise continuous in general, in the topology of the real line,
but all observable predictions of the theory would still remain unchanged,
despite the mathematical monstrosity and the absence of any simple relation-
ship between the sub-carrier space L and physical space SF.
I believe that this is, in the more precise language of modern algebra, the
expression of Poincare's idea about the conventionality of space.
The author acknowledges interesting discussions on the subject with P.
Benioff, R. Geroch, R. Haag, P. Havas, M. A. Melvin, and 1. Segal.

Centre de Physique Theorique, CNRS,


Marseille, Luminy, France

NOTE

1 This invariant pseudo-distance is unique up to a factor.


74 HANS EKSTEIN

REFERENCES

[1) Poincare, H., 1962. Science et Hypothese. New York: Dover.


[2) Griinbaum, A., 1973. Philosophical Problems of Space and Time. Second, en-
larged edition. Boston Studies in the Philosophy of Science, Vol. 12. Dordrecht,
Holland: D. Reidel.
[3) Ehlers, J., 1973. 'The Nature and Structure of Spacetime.' In The Physicist's Con-
ception of Nature, ed. by J. Mehra, pp. 71-91. Oordrecht, Holland: D. Reidel.
[4) Avishai, Y., and H. Ekstein, 1974. 'Einstein~s Equivalence Principle. An Explicit
Statement and Its Derivation from Special Relativistic Presymmetry.' Commun.
Math. Phys. 37,193.
[5) Oeser, S., 1970. 'Self-Interaction and Gauge Invariance.' General Relativity and
Gravitlltion 1,9.
[6) Thirring, W., 1961. 'An Alternative Approach to the Theory of Gravitation.' Ann.
Phys. 16,96.
[7] Schild, A., 1962. 'Gravitational Theories of the Whitehead Type and the Principle
of Equivalence.' In Evidence for Gravitational Theories, ed. by C. Millner, pp. 69-
115. New York: Academic Press.
[S] Einstein, A., 1973'. 'Reply to Criticisms.' In Albert Einstein, Philosopher Scientist,
ed. by P. SchUpp, p. 677. La Salle, lll.: Open Court Publ. Co.
[9] Ue,S.,1935. Ges. Abhandlungen, VoL 2/1, pp. 374-469. Leipzig: B. G. Teubner.
[10] Mackey, G. W., 1963. Mathematical Foundations of Quantum Mechanics. New
York: Benjamin.
[11) Havas, P., 1964. 'Four-Dimensional Formulations of Newtonian Mechanics and
Their Relation to the Special and General Theory of Relativity.' Rev. Mod. Phys.
36,93S.
[12) Weinberg, S. 1972. Gravitation and Cosmology. New York: J. Wiley.
CATHERINE Z. ELGIN

THEORY REDUCTION:
A QUESTION OF FACT OR A QUESTION OF VALUE?

Frequently the reduction of scientific theories is treated as a linguistic issue:


One theory is said to be reduced to another if the objects referred to by the
former are identified with entities in the domain of the latter, and the laws of
the former are derived from the laws of the latter (plus whatever connecting
principles, correspondence rules, or bridge laws are needed to link the
vocabularies of the two theories). The objects of the reduced theory are thus
shown to be nothing but objects (or combinations of objects) recognized by
the reducing theory, and the concepts of the reduced theory are shown to be
theoretically superfluous. The resulting ontological and conceptual economies
may reasonably be construed semantically, for they demonstrate that the
language of science requires fewer primitive terms than had been previously
supposed. Positions taken by Quine and Goodman suggest, however, that
evidential and linguistic arguments are in principle too weak to secure theore-
tical reduction.
The failure to discover a syntactic or semantic solution to Goodman's
paradox suggests that there are no formal criteria for determining which
generalizations are lawlike. Let 'x is grue' be defined as 'x is green if examined
before time t; otherwise x is blue'. There appear to be no formal arguments to
demonstrate that 'All emeralds are green' is lawlike, but 'All emeralds are
grue' is not. Why then should we accept the former rather than the latter
since (t being some future time) both are equally compatible with the
evidence? This is the new riddle of induction.
Quine argues that language must be conceived holistically. With the
repudiation of the analytic/synthetic distinction we lose any objective ground
for separating out favoured usages as giving the meaning of a term while
taking others to convey contingent beliefs about its objects. Indeterminacy of
translation and ontological relativity are consequences of the holism of
language. So long as dispositions to verbal behavior are preserved, there is no
factual basis for preferring one translation manual to another. And we
individuate objects by means of a system of linguistic devices - "plural
endings, pronouns, numerals, the 'is' of identity, and its adaptations 'same'
and 'other,."l Thus the reference of our terms and the extensions of our
predicates are relative to the interpretation of those devices. There is then no

75

R. S. Cohen and M. W. Warto/sky (eds.), Physical Sciences and History o/Physics, 75-92.
© 1984 by D. Reidel Publishing Company.
76 CATHERINE Z. ELGIN

saying absolutely what the terms of a language refer to or whether two


languages refer to the same things.
If syntactic and semantic arguments are incapable of determining which
generalizations are lawlike, and incapable of deciding in any absolute sense
what the objects of a theory are, they are (to put it mildly) unlikely to be
sufficient to demonstrate that the objects of one theory are identical to
(some of) those of another, and that the laws of the one can be derived from
the laws of the other. Accordingly, even if the result of theoretical reduction
is a change in the syntax and semantics of the language of science, it should
not be thought that only linguistic and evidential arguments are required to
achieve this result. Or so I will argue.
The structure of my paper is as follows: I begin by considering Davidson's
argument that anomalous monism yields a solution to the mind/body
problem. Anomalous monism is the thesis that every mental event is physical,
but it is not the case that mental events can be given purely physical explana-
tions. Davidson recognizes the validity of Quine's arguments. But he claims
that the formal question of lawlikeness - that is, whether a generalization, if
confirmed by the evidence, is to be taken as lawlike - is a question of the 'fit'
between the predicates in the generalization. "Nomological statements bring
together predicates we know a priori are made for each other." 2 Thus on his
account 'All emeralds are grue' is not lawlike because 'emerald' and 'grue' are
not suited to each other. Correspondingly, he denies that psychophysical
generalizations are lawlike because mental and physical predicates are not
suited to one another. And he argues that the indeterminacy of translation
demonstrates that psychological and psychophysical generalizations cannot
be reduced to purely physical laws: There are indefmitely many mappings
of them onto our physical theory and there is no fact of the matter as to
which one is correct. Although he accepts the thesis that every mental event
is in fact identical with some physical event, Davidson concludes on the
basis of linguistic arguments that there are no laws linking events described
as mental with events described as physical. In effect, he believes that the
mind/body problem can be solved by the philosophy oflanguage.
I argue that the linguistiC arguments he adduces are not sufficient for his
conclusion (or if they are, then they constitute a general a priori argument
against theoretical reduction). Nevertheless, Davidson's account is important
for at least two reasons: First it demonstrates the tenability of a token-token
identity thesis. It may be the case that every object in the domain of one
theory is identical with an object (or combination of objects) in the domain
of another, without its being the case that the kinds recognized by the former
THEORY REDUCTION 77

are identical with kinds recognized by the latter, or that laws expressed in
terms of the conceptual apparatus of the former are reducible to laws
expressed in terms of the conceptual apparatus of the latter. Second,
although his appeals to language are not themselves sufficient, the intuitions
that motivate them may have as their basis sound reasons for blocking reduc-
tion. Davidson's account suggests that the different disciplines may have con-
ceptual commitments that preclude reduction even when the necessary corre-
lations can be established.
I hope to show that the decision as to what generalizations ought to be
treated as lawlike and the decision as to whether the objects of one theory
ought to be identified with the objects of another are normative decisions.
They are based on the interests and cognitive values of the different disci-
plines, and perhaps also on more general values that those disciplines are
designed to serve. If the disciplines differ in the values that they recognize,
theory reduction may be blocked even though the evidence supports corre-
lations between generalizations warranted by the two theories.
My discussion focuses on the case of .psychophysical reduction, both
because previous discussions of token-token identity have done so, and
because it is fairly easy to show that normative issues are relevant to deciding
this case. I think that these issues arise for other proposed theoretical reduc-
tions as well - even for the celebrated reduction of chemistry to physics. But
since the normative dimensions of our activities are typically brought to our
attention by disagreements, if our investigation of reduction is restricted to
disciplines that share cognitiye values, the fact that there is a normative aspect
to the problem is likely to be overlooked.
Davidson draws the distinction between the mental and the physical lin-
guistically: A term is mental if and only if it is intentional; otherwise it is
physical (p. 84). Accordingly, an event is mental if and only if it uniquely
satisfies an open sentence that employs at least one mental term essentially.
A mental event is thus the object of a propositional attitude. An event is
physical if and only if it uniquely satisfies an open sentence that employs
only physical terms essentially. It follows that the same event can be both
mental and physical.
Davidson's argument rests not on the de facto failure of psychology to
discover rigorous, exceptionless, predictive laws, but rather on the conviction
that psychological and psychophysical generalizations are not fully lawlike.
Determining under what circumstances a generalization is lawlike is thus of
central importance for Davidson's philosophy of mind.
Generalizations are lawlike which are supported by their instances and
78 CATHERINE Z. ELGIN

sustain counterfactuals (p. 92). Since evidential support is never strong


enough to require us to recognize a generalization as lawlike, "ruling it
lawlike must be a priori" (p. 90). Like his criterion of the mental, Davidson's
criterion of lawlikeness is linguistic. His claims that "lawlikeness is a matter
of degree" (p. 92), and argues that the degree of lawlikeness of a generaliza-
tion is a function of the fit between its terms. He does not explain what it
is for predicates to be "made for each other" or "suited to one another"
(p. 92). like the lawlikeness of the generalizations that result, the compati-
bility of predicates is a matter of degree. What is unclear is the nature of the
a priori linguistic knowledge which is supposed to determine the intimacy of
the relation between predicates of a given generalization, and hence the
degree of inductive support its instances can afford.
Although "mental and physical predicates are not made for one another"
(p. 93), Davidson claims that psychophysical generalizations are more lawlike
than 'All emeralds are grue.' We apparently know a priori that psychophysical
claims are well formed and can be confirmed by evidence. Such claims in
turn confirm rough psychophysical generalizations which are lawlike to the
extent that they provide "good reason to expect other cases to follow suit
roughly in proportion" (p. 93). Although they are "assumed to be only
roughly true, or ... are explicitly stated in probabilistic terms, or ... are
insulated from counterexample by generous escape clauses" (p. 93), these
generalizations (like those in Hempel's explanation sketches) can be refined
to the point where they state precise, exceptionless, predictive laws. The
critical issue is what form such refinement takes - specifically, in what voca-
bulary must such refinement be expressed?

On the one hand, there are generalizations whose positive instances give us reason to
believe the generalization could be improved by adding further provisos and conditions
stated in the same general vocabulary as the original generalization. Such a generalization
points to the form and vocabulary of the finished law: we may say that it is a homo-
nomic generalization. On the other hand there are generalizations which when instan-
tiated may give us reason to believe that there is a precise law at work, but one that can
be stated only by shifting to a different vocabulary. We may call such generalizations
heteronomic (p. 94).

A homonomic generalization is one whose vocabulary is that of a compre-


hensive closed theory. Unless a theory is comprehensive and closed, events
within its domain causally interact with events outside of its domain. Since
the latter are not describable in the vocabulary of the theory, such causal
sequences are not subsumable under the laws of the theory. In that case the
THEORY REDUCTION 79

theory is unable to explain the occurrence of every event within its domain.
Davidson contends that only homonomic generalizations are fully lawlike, for
heteronomic generalizations, by their very form, require replacement rather
than revision to yield explicit, precise, exceptionless laws.
Davidson's argument for the anomalism of the mental rests essentially on
the heteronomic character of psychophysical generalizations. lIDs seems a
bit strange, for the question one wants to ask is whether mental events are
or are not governed by explicit, predictive, exceptionless laws. And this
question at least seems to be independent of questions concerning the seman-
tic character of our descriptions and generalizations regarding such events.
Davidson contends, however, that such independence is illusory. Mental
events are identified as the objects of propositional attitudes. Indeed, "events
are mental only as described" in the vocabulary of propositional attitudes
(p. 89). Hence, if terms belonging to that vocabulary are incapable of entering
into statements of genuine laws, events identified as mental are anomalous.
Because of the description-relative character of the mental, any attempt at
redeSCription is precluded. "[T] 0 allow the possibility of ... laws [linking
the mental and the physical] would amount to changing the subject ...
deciding not to accept the criterion of the mental in terms of propositional
attitudes" (p. 90).
Propositional attitudes constitute a network and the content of each pro-
positional attitude depends on its place in that network. Accordingly, a
mental term derives its meaning from its place in the descriptive system
representing the network of propositional attitudes. "There is no assigning
beliefs to a person one by one on the basis of his verbal behavior, his choices,
or other local signs no matter how plain and evident, for we make sense of
particular beliefs only as they cohere with other beliefs, with preferences,
with intentions, hopes, fears, expectations and the rest" (p. 96). Davidson's
point is not epistemic, but ontic - not that we don't know how to assign
beliefs one by one, but rather that the very identity of a belief depends on
its coherence with other beliefs, preferences, and so on. It makes no sense to
be an atomist about propOSitional attitudes.
Davidson admits that physical theory and the network of propositional
attitudes determine descriptive systems capable of representing a common
range of events. Despite their common range, however, he argues that the
disparate commitments of the physical and the mental reveal the systems
to be genuinely distinct.

It is a feature of physical reality that physical change can be explained by laws that
80 CATHERINE Z. ELGIN

connect it with other changes and conditions physically described. It is a feature of the
mental that the attribution of mental phenomena must be responsible to the background
of reasons, beliefs, and intentions of the individual. There cannot be tight connections
between the realms if each is to retain its allegiance to its proper source of evidence
(pp.97-98).

Evidence concerning physical theory must be described in physical terms.


Only then can the relation of an event to physical laws and to other events
in its physical environment be shown; only under such a deSCription can it
enter into a physical explanation of the course of events. Conceived or
described in any other way, that event would be evidence neither for nor
against the theor~ in question. Correspondingly, evidence concerning the
mental realm must be described in terms of propositional attitudes. Only then
can its relation to background beliefs, intentions, and reasons be exhibited;
only then can it enter into psychological or intentional explanations of the
course of events. Even though the vocabularies proper to the mental and the
physical refer to the same events, the disparate evidential and explanatory
commitments of the two systems require that they provide distinct con-
ceptions of those events.
By itselfthe argument that terms referring to mental events have systemati-
cally different criteria of application from terms referring to physical events
does not entail the absence of a lawful connection between the two realms.
Davidson contends, however, that it does so when conjoined with the princi-
ple of charity: We must impute coherence to the beliefs, desires, intentions
and actions that we attribute to an agent. "[W] hen we use the concepts of
belief, desire, and the rest, we must stand prepared, as the evidence accumu-
lates, to adjust our theory in the light of considerations of overall cogency:
the constitutive ideal of rationality partly controls each phase in the evolu-
tion of what must be an evolving theory" (p. 98). The attribution of
rationality, and hence of a coherent system of beliefs, desires, preferences,
and the like, is required to treat persons as persons. Davidson's position then
is that explanations in terms of propositional attitudes are subject to a
methodological constraint that is lacking in explanations in terms of physical
characteristics. It is this methodological constraint, which stems from the
concept of a person, that precludes a lawful relation between the mental and
the physical.
The proper interpretation of the principle of charity is critical. If no
evidence could override the imputation of rationality, then the anomalism
of the mental would surely follow. For whenever incoherence threatened we
would be required to make ad hoc re~djustments in our attribution of propo-
THEORY REDUCTION 81

sitional attitudes. And since the content of each propositional attitude


derives from its place in the system, the required readjustments would con-
situte subtle modifications in the criteria for the application of mental
terms. Rather than recognize irrationality we would be required to revise not
only our interpretation of the agent's beliefs, desires, and meanings, but
also our understanding of what belief, desire, and meaning are. The dynamic
character of mental types would obviously preclude their identification with
static physical types.
But of course under certain circumstances we do call actions, beliefs,
preferences, and desires irrational. And this makes the interpretation and
status of the principle of charity problematic. To admit that it is some-
times reasonable to attribute an incoherent system of propositional attitudes
to an individual is not necessarily to introduce global confusion into the
realm of the mental. So long as we have a "background of true [or warranted]
belief against which ... failure can be construed" (pp. 96-97), it makes
sense to recognize some degree of irrationality and error. One might argue,
therefore, on a more charitable reading, Davidson's principle would not
require that the criteria for the application of mental terms be fluid enough
to be modified each time any individual's system threatened to become
incoherent. Rather, it would require that our concepts of belief, preference,
intention and the like be such that systems of propositional attitudes
generally turn out to be reasonably coherent and rational and that, ceteris
paribus, those concepts are to be applied in such a way as to minimize the
incoherence and irrationality of each individual's set of propositional
attitudes.
On this interpretation, the principle of charity seems to be simply a
consequence of the holism of the mental. Since the content of each pro-
positional attitude depends on its place within a system, without a reasonably
coherent system we would be unable to identify particular propositional
attitudes. And since a major motive for introducing propositional attitudes
in the first place is to give explanations for actions in terms of reasons, with-
out the imputation of rationality, we would fail at our assigned task. Thus to
the extent that systems of propositional attitudes deviate from coherence and
rationality, our faith in the particular identifications should be diminished.
Some such principle, however, is necessary to any holistic account. The
introduction of a hypothetical entity or system of hypothetical entities
is unjustified unless that entity or system actually serves the purpose for
which it was introduced - unless it explains what it was introduced to
explain. And whenever entities are identified on the basis of their place in a
82 CATHERINE Z. ELGIN

system, our knowledge of their existence and character is surely conditional


on the coherence and explanatory power of the system. Accordingly, the
claim that the principle of charity is a methodological rule peculiar to the
realm of the mental appears unfounded. It is but an instance of the general
(and uncontroversial) methodological requirement that hypothetical entities
serve the explanatory purposes for which they are introduced.
Davidson's discussion of the holism of the mental does, however, force-
fully argue that propositional attitudes must be conceived as hypothetical
entities. Mental terms, then, designate hypothetical entities. Their introduc-
tion as part of an explanatory scheme is vindicated, and the existence of their
referents (defeasibly) demonstrated if they lead to good explanations and
predictions of behavior. The success of the psychophysical generalizations
made possible by their introduction is thus critically important. The
systematic failure of defmitional behaviorism demonstrates that there is no
simple correlation between mental and behavioral predicates (p. 92). Since an
entire system of propositional attitudes must be postulated to account for
intentional action - indeed, even to identify a given event as an intentional
action - explanations employing mental terms cannot reasonably be con-
strued as abbreviated descriptions of behavior. Rather, mental terms identify
entities that are thought to be responsible for the production of certain
behaviors, and it is because these behaviors can be accounted for by reference
to the interrelation of beliefs, desires, preferences, and the like that they are
identified as intentional actions.
According to Davidson, our current theory of the mental is heteronomic
because events described in terms of propositional attitudes have causes or
effects that are described without the use of mental terms. He concludes that
to arrive at precise, exceptionless laws, our mental discourse would have
to be replaced by (rather than just augmented by) the vocabulary of a com-
prehensive closed physical theory. The precise laws that underlie our psycho-
logical and psychophysical generalizations cannot be expressed in the same
general vocabulary as those generalizations.
It is difficult to know how to evaluate this claim. Davidson does not
explain how to determine whether proposed modifications belong to the
same general vocabulary as the original generalization. Since his aim is to dis-
tinguish between the vocabulary of propositional attitudes and that of
physical theory, one might suppose he means to employ the distinction
between intensional and extensional discourse. That distinction, however, will
not serve his purpose. Since laws, on Davidson's account, sustain counter-
factual and subjunctive claims, they (and the explanations in which they
THEORY REDUCTION 83

occur) are intensional. Davidson gives no reason to suppose that homonomic


generalizations and generalizations employing the vocabulary of proposi-
tional attitudes differ in logical form. Therefore his argument does not
demonstrate that differences between the vocabulary of a homonomic genera-
lization and that of a heteronomic generalization can be explicated by
reference to the logic of the explanations in which they occur.
In any case, Davidson does take the question of psychophysical reduction
to be a question of the relation between two distinct vocabularies: the
vocabulary of the mental and the vocabulary of the physical. Because the
distinction between heteronomic and homonomic generalizations is a
distinction in the form that theoretical development will take, the growth of
science is of central concern. Theoretical development involves the identifica-
tion of entities described in terms dictated by one theory with entities
described in terms dictated by another. Since such identifications are
described by correspondence rules, the proper interpretation of correspon-
dence rules is crucial.
In order for our psychophysical generalizations to lend support to Singular
causal claims and related explanations of particular events, as Davidson
maintains that they do, the underlying laws must explain the success of those
generalizations. Traditionally, such generalizations would be treated as
empirical laws whose success is explained by their derivation from the under-
lying theory. Oearly such a procedure is unavailable to Davidson. For if it
were derivable from theoretical laws, a generalization would be homonomic; it
could not contain predicates alien to the laws from which it was derived.
Further, any such derivation would establish a type-type identity; events of
the kind mentioned in the psychophysical generalization would be shown
to be identical with events of the kind picked out by the physical predicates
of the theoretical laws.
Sellars offers an alternative account of the success of nontheoretical
(or relatively nontheoretical) generalizations.3 He argues that the underlying
theoretical laws explain the facts that there are, and hence derivatively
explain why those facts conform to the generalizations to the extent that
they do. On the surface Sellars's account seems ideally suited to explain the
success of psychophysical generalizations. But it cannot do so within the
framework of Davidson's program. The facts and events that a theory
explains are facts and events as described in the vocabulary of that theory.
In order to explain the success of nontheoretical generalizations, there must
be correspondence rules which bridge the gap between nontheoretical and
theoretical vocabularies. But such correspondence rules render the original
84 CATHERINE Z. ELGIN

generalizations homonomic by bringing their predicates into the vocabulary


of the theory. Thus the establishment of appropriate correspondence rules
would render psychophysical generalizations homonomic. Indeed, at least
where there were precise, true, psychophysical generalizations, such
correspondence rules would establish the identity of types of mental events
with types of physical events. It follows that it is a requirement of Davidson's
program that correspondence rules are not treated as part of the closed
comprehensive physical theory that explains the occurrence of events having
mental descriptions. Without correspondence rules, however, the underlying
physical theory only explains the occurrence of causal sequences which in
fact have psychOl>hysical descriptions; it cannot explain their having the
psychophysical descriptions that they do. But then the underlying physical
theory cannot explain the success of our psychophysical generalizations; it
will not show that those generalizations genuinely support singular causal
claims and explanations of particular events.
These considerations are not restricted to the case of psychophYSical
generalizations. If Davidson is correct in distinguishing between homonomic
and heteronomic generalizations, then because correspondence rules would
render any domain homonomic, they cannot belong to any comprehensive
closed theory. Therefore, the explanatory success of any truly heteronomic
generalization cannot be explained by the underlying homonomic theory.
Our faith that most of the generalizations that make up science and practical
wisdom lend support to singular claims and to the explanation of particular
events cannot, on the view that Davidson is committed to, be supported.
Davidson takes the system of correspondence rules linking two theories to
constitute a translation manual - a set of rules for translating from the
vocabulary of one theory into the vocabulary of the other. He contends that
it follows from the indeterminacy of translation that the reduced theory
might be mapped onto the reducing theory in a variety of ways, each com-
patible with the total evidence. He concludes that there is no fact of the
matter regarding which mapping establishes the reduction, no fact of the
matter regarding which set of correspondence rules is correct.
The indeterminacy of translation thus appears to entail the anomalism of
the mental. Because the system of propositional attitudes is an open system -
because not every event that is the cause or effect of an event describable in
the vocabulary of propositional attitudes is itself describable in that voca-
bulary - it must be extended or reduced to yield a comprehensive closed
theory. But any attempted extension or reduction runs afoul of the problem
of the indeterminacy of translation. Reduction of the network of proposi-
THEORY REDUCTION 85

tional attitudes to neurophysiological theory, for example, requires that the


events described as believings, intendings, desirings, etc. be identified with the
entities recognized by that theory. Since there are potentially a variety of
acceptable systems of correspondence rules,each yielding a different set of
identifications, the selection of any one such system is ontologically arbitrary.
Although considerations such as simpliCity, intuitive appeal, ease of calcula-
tion, etc. might influence the choice of a system, there are no factual or
ontological considerations that determine a uniquely correct choice.
The same problem arises when we attempt to expand or refine a theory. In
the one case equivalence principles are required to establish that the entities
described as F's in the restricted theory are among the things (or are the very
same things) we describe as F's in the extended theory. In the other, such
principles are required to show that what had formerly been described as F's
are considered to be G's in the refined theory. Theory expansion or refine-
ment, then, like theory reduction, is subject to the indeterminacy of transla-
tion.
The foregoing discussion demonstrates, however, that the problem of
radical translation cannot justify treating some developing theories as
homonornic and others as heteronomic. If a system of correspondence rules
constitutes a translation manual, then the relation between a theory and its
successors is always mediated by analytical hypotheses. Hence, if radical
translation is to be our guide, developing theories are all heteronomic.
Indeed the problems raised by this account are not restricted to developing
theories. Because different sciences and different levels of complexity within
a science are related by correspondence rules, if Davidson's account is correct,
those relations render even a finished science heteronomic. Using Davidson's
criterion we are forced to recognize the anomalism of the chemical because
we cannot establish a unique reduction of chemistry to physics, and even the
anomalism of the physical, because there is more than one mapping of
physical theory onto itself.
The anomallsm of the mental is then a consequence of neither the special
character of the mental nor the relation between reducing and reduced
theories in science. Rather, it rests on the fact that whenever we attempt to
establish that two descriptions are descriptions of the same thing, our
accounts are vitiated by the inscrutability of reference and the resulting
indeterminacy of translation. There is no fact of the matter with respect to
which our identifications are correct or incorrect.
Indeterminacy, then, is a universal problem, infecting not only translation
between radically disparate languages, but also ordinary discourse within a
86 CATHERINE Z. ELGIN

single language, and even the monologues of each individual speaker. But does
indeterminacy raise special problems for theoretical reduction beyond those
infecting all oflanguage? I think not.
Unlike ordinary discourse, scientific theories which are candidates for re-
duction are well regimented. The criteria for the individuation of the objects
that a theory recognizes - indeed, its entire referential apparatus - are set
forth explicitly when the theory is expressed in canonical notation. If both
the reducing and the reduced theory are cast in canonical notation, making
their ontological commitments explicit (a requirement for the establishment
of correspondence rwes in any case), then there is a fact of the matter concern-
ing the identification of the entities recognized by the reduced theory with
those recognized by the reducing theory. If, for example, chemical theory
provides criteria for the individuation of chemical states and physical theory
provides criteria for the individuation of physical states, then with which
physical state(s) a given chemical state is to be identified is a question of fact.
Correspondingly, if neurophysiology came to yield criteria for the individua-
tion of brain states, and psychology came to yield criteria for the individuation
of propositional attitudes, there would be a fact of the matter regarding which,
if any, propositional attitudes were to be identified with particular brain states.
Relations between reducing and reduced theories are not, as Davidson
believes, established by means of a translation manual or system of analytical
hypotheses. Rather, correspondence rules are synthetic claims of a global
theory or conceptual scheme that comprehends both the reducing and the
reduced theory. According to this ac.count then, correspondence rules have
truth values.
The fact that the objects of one theory are identical with objects (or com-
binations of objects) of another, and that the predicates of the first are co-ex-
tensive with predicates of the second is not sufficient to demonstrate that the
former theory is reducible to the latter. For reduction to be justified, the
two theories must be lawfully connected. The question then is this: Given
that we have true (or, at any rate, warranted) correspondence rules linking
the two theories, what more is required to recognize them as lawlike?
Recall that when one theory is successfully reduced to another the con-
ceptual apparatus of the reduced theory is shown to be superfluous.
Accordingly, that conceptual apparatus can in principle be eliminated with no
loss to science. At least two questions must be answered in order to
determine whether one theory can be eliminated in favor of another with no
loss to science: The first concerns the growth of science; the second, the
standards of adequacy for scientific explanations.
THEORY REDUCTION 87

The question whether two theories make reference to the same objects and
characterize them by means of co-extensive predicates has a determinate
answer only if theories are construed as systems of sentences. Reduction then
is always reduction of theories as regimented by a canonical notation. But
both the reducing theory and the reduced theory belong to growing,
developing sciences. And as they grow and develop, the structure of their
theories - as exhibited in their canonical forms - changes. If their common
subject matter contains recalcitrant evidence, the theories must be modified
to accomodate it.
Quine and Duhem have demonstrated that there is no unique correction
dictated by either the structure of the theory or the structure of the evidence.
If the two theories retain conceptual or methodological autonomy, then
within the context of their respective research programs, alternative types of
modification might be appropriate. For a variety of reasons, stemming from
each theory's conception of its own domain, its emphasis, its methodology,
its characteristic problems, and its values, it might be rational to revise them
in different ways. The scientific unity achieved by reduction is but one
value that must be weighed against others in decisions regarding the rational
preferability of different modifications.
Accordingly, although it is based on the relation between the sets of
sentences that constitute our current theories, reduction has a prospective
and regulative aspect as well. Since the result of reduction is the elimina-
bility of the conceptual apparatus of the reduced theory, the recognition of
correspondence rules as lawlike involves a commitment to the direction of
scientific progress: the commitment that henceforth both sciences are to
agree as to which corrections are to be made in order to bring their common
theory into accord with the evidence. With the elimination of the conceptual
resources of the reduced theory, it~ basis for autonomous development is
lost to science. It follows that in deciding whether the conceptual resources
of one theory can be eliminated in favor of those of another, our attention
cannot be restricted to the current theories. We must consider the prospects
for scientific development as well.
At least part of what is involved in claiming that a theory can be
eliminated in favor of another with no loss to science is that the explanations
of the former can be replaced by explanations of the latter without loss. This
clearly requires that the standards for adequate explanations be shared by the
two theories. If this is not the case, then even if we demonstrate that the objects
of the two theories are identical and the predicates coextensive, it is not clear
that the correspondence between the two should be taken as lawlike.
88 CATHERINE Z. ELGIN

If the standard Hempelian account yields sufficient conditions for


adequate explanations, then this raises no problems. But if explanation is
interest-relative, and if two theories that treat of the same objects can yield
explanations that serve different interests, then the reduction of one to the
other will involve a loss of explanatory power. Although to each explanation
of the reduced theory there will correspond an explanation of the reducing
theory, the latter will not be adequate to the purposes that the original
explanations were designed to serve.
For example, it is reasonable to suppose that biology will recognize as
laws certain generalizations concerning the goal-directed behavior of animals.
According to Davidson, however, to treat human beings as pers~s is to
rationalize their behavior, and this involves treating their actions as
anomalous. Accordingly, if Davidson's claim is correct, generalizations that
function as laws in biology cannot be recognized as laws by the social
sciences. Their status as law is suspended when the focus of interest is
restricted to man.
It is not because we believe homo sapiens to constitute a singularity in
the natural order that we refuse to take such biological generalizations as
the basis for our explanations of human action. Rather it is because we
recognize that our explanations of human action are subject to a requirement
- that of rationalizing behavior - to which our accounts of the behavior of
other animals are not. This is not to deny that there are contexts of inquiry in
which it is appropriate to treat human goal-directed behavior on a par with
the goal-directed behavior of other animals. But where our purpose is to
rationalize behavior - as it is in the social sciences and generally in explana-
tions in which human behavior is conceived as action - the true, warranted,
universal generalizations in question are not to be recognized as laws.
Davidson explains the supervenience of the mental on the physical by
saying that ''there cannot be two events alike in all physical respects but
differing in some mental respect, or that an object cannot alter in some
mental respect without altering in some physical respect" (p. 88). If my
account is correct, then the supervenience of the mental on the physical
amounts to this: Although every mental event is physical, and all of its
characteristics are physical characteristics, it is not the case that every
adequate explanation of it as physical is an adequate explanation of it as
mental. And in general science A is supervenient on science B if the objects
recognized by A are in the domain of B, but the standards for adequate
explanations dictated by A are not shared by B.
One might reasonably accept my claim that there is a fact of the matter
THEORY REDUCTION 89

regarding the identification of objects of theories cast in canonical notation,


but deny its relevance to the psychophysical case. The system of proposi-
tional attitudes is so little understood that we just don't know what a
regimented psychological theory would be like. Such an objection is, of
course, sound. But it does not represent Davidson's position. "The thesis is
... that the mental is nomologically irreducible: there may be true general
statements relating the mental and the physical, statements that have the
logical form of a law; but they are not lawlike ... if by absurdly remote
chance we were to stumble on a nonstochastic true psychophysical generaliza-
tion, we would have no reason to believe it more than roughly true" (p. 90).
Note what is being claimed here: Although psychophysical generalizations
may be true, none are lawlike. They must therefore be regarded as accidental
truths - truths which establish de [acto correlations between the mental and
the physical, but which are logically too weak to support reduction. Further,
were we to consider a true psychophysical generalization, we would never
really believe it. Regardless of the weight of evidence supporting it, we would
take it to be only approximately true.
Davidson claims that to treat an individual as a person we must treat him
as rational (p. 98), and that to treat him as rational, we must describe his
behavior in the vocabulary of propositional attitudes. He concludes that to
admit the possibility of psychophysical laws amounts to changing the subject
- "deciding not to accept the criterion of the mental in terms of the vocabu-
lary of propositional attitudes" (p. 90). That is, deciding not to interpret
that behavior as the action of a person. Because the reduction of one theory
to another justifies the elimination of the vocabulary of the former from the
language of science, the reduction of the mental to the physical would justify
the elimination of the vocabulary of propositional attitudes. By ceasing to
identify mental events in the way that we used to, we change the subject.
The replacement of a given set of criteria by a more accurate or more
useful set is, of course, a normal feature of scientific development. If this is
changing the subject, then science changes its subject regularly. If the psycho-
physical case is an exception to this general practice, it is because the
importance of the concept of a person makes the subject one we are parti-
cularly reluctant to change.
Davidson is, in effect, stipulating what it is to be a person. Stipulating
criteria for membership in a kind is a common and (in general) sound seman-
tic procedure. There is an aspect of this particular stipulation, however, that
should be noted. Normally when a stipulation establishes criteria for
inclusion in a kind, it is a question of fact whether anything satisfies those
90 CATHERINE Z. ELGIN

criteria. Since it is a question of fact, the methods of science are to be


employed to discover the answer. Davidson contends, however, that science
can tell us nothing about what individuals (if any) satisfy the criteria for.
being a person. Science is thus incapable of sustaining or overruling the
employment of those criteria. Descriptions of behavoir in terms of propo-
sitional attitudes are indefeasible by any possible scientific findings.
Davidson's position is this: Even if we had a comprehensive closed physical
theory and true psychophysical generalizations linking the system of propo-
sitional attitudes to that theory, to treat those generalizations as lawlike
would be to give up treating individuals as persons.
Can such a position be justified? Davidson's purely semantic account
cannot do so. It can, to be sure, stipulate criteria for the application Of the
term 'person' which make his position coherent. Since those criteria pre-
suppose the anomalism of the mental, they guarantee that recognizing
psychophysical laws amounts to changing the subject. But such an account
cannot explain why we should go to such extremes to avoid changing the
subject in question.
Epistemologically the situation is this: Since Hume it has been recognized
that we are never epistemically obliged to draw a general conclusion from a
finite base of evidence or to treat a generalization as lawlike. Choices between
options are the result of two factors: the relative values of different outcomes
and the probability of those outcomes. Since a finite body of evidence never
gives a general conclusion a probability of 1, the decision to accept a general
claim on the basis of instances, and the decision to treat it as lawlike are made
under uncertainty and risk. In deciding whether to recognize psychophysical
generalizations as lawlike, we must weigh the benefits of succeeding in
reducing the mental to the physical against the costs of falsely believing we
have done so. If we refuse to treat psychophysical generalizations (however
well confirmed) as lawlike, it is because we recognize that no matter how well
confirmed our theory is, it still might be false, and we estimate that the cost
of mistakenly adopting a type-type identity theory would far outweigh the
benefits to be gained from successful psychophysical reduction.
The weighting is, and I think must be, ideological. Central to the ideology
underlying our moral and political theory is the view of a person as a rational,
responsible agent. The notions of rationality and responsibility (obscure
though they are) involve a recognition of the right of the individual to con-
ceive of his behavior (and to have others conceive of it) as consisting of
actions undertaken on the basis of his own beliefs, desires and preferences.
To accept a type-type identity theory is to accept the view that each
THEORY REDUCTION 91

individual's behavior is determined by psychophysical laws. This may indeed


be true. But if it is false, and we do not know how to reconcile free will with
determinism, then in accepting it we commit the grave injustice of treating
human bein~ as deterministic automata rather than as responsible agents.
And it is the recognition of the magnitude of that potential injustice that
makes us reluctant to accept psychophysical reduction. Given this evaluation
of the seriousness of the risk, we would be not only morally, but also epis-
temically irresponsible to take it.
Our epistemic goals are achieving knowledge and avoiding error. Where the
two diverge, we must choose between them. If we deny that warranted
psychophysical generalizations are lawlike, our primary concern is to avoid a
particular error - the error of falsely believing the behavior of human beings
to be fully determined by psychophysical laws and therefore wrongly denying
that we are rational, responsible persons. In our choice whether to treat
psychophysical generalizations as lawlike, the value of a unified scientific
theory conflicts with ideological values. In this case, ideological values prevail.
Although I do not have time to argue it now, I want to suggest that the
features of psychophysical reduction that I have emphasized are characteristic
of reduction in general. The reduction of one theory to another is straight-
forward only if, in addition to satisfying the evidential and semantic require-
ments, the reducing and the reduced theories share cognitive values. Where
one theory is supervenient on another, we cannot maintain that the one is
reducible to the other without loss of explanatory power. For there is no
guarantee that explanations acceptable from the point of view of the reducing
theory are also acceptable from the point of view of the reduced theory, or
that the interests served by the former will be the same as those served by the
latter. It will, to be sure, sometimes happen that the loss in explanatory
power is outweighed by the benefits of a unified theory. In these instances
reduction should be carried through. But it requires a normative argument to
demonstrate that this is (or is not) the case.
In arguing that there is an ineliminably normative aspect to theory
reduction, I do not mean to suggest that this aspect is subjective, or idiosyn-
cratic, or arbitrary. In fact it was precisely to avoid having to decide arbi-
trarily whether or not correspondence rules are lawlike that I was driven to
appreciate the role of norms. It is clear that the decision to recognize one
theory as reduced to another (and therefore eliminable in favor of the other)
must be grounded in good reasons if it is to be scientifically valid. If my
account of reduction is correct, it is incumbent on the philosophy of science
to discover what constitute good reasons for saying that a generalization
92 CATHERINE Z. ELGIN

ought to be regarded as lawlike. The arguments of Quine and Goodman show


that such facts of the matter as there are insufficient to decide the case. I
suggest that to solve the problem we must look also to the interests that the
science is designed to serve, the sorts of explanations it intends to provide,
the range of questions it seeks to answer, and the cognitive values it is com-
mitted to uphold.

University ofNorth Carolina

NOTES

1 W. V. Quine, 'Ontological Relativity,' in Ontological Relativity and Other Essays (New


York: Columbia University Press, 1969), p. 32.
2 Donald Davidson, 'Mental Events,' in Experience and Theory, ed. Lawrence Foster
and J. W. Swanson (Amherst: University of Massachusetts Press, 1970), p. 93. Page
numbers in parentheses throughout my paper refer to this article.
3 Wilfrid Sellars, 'The Language of Theories,' in Science, Perception, and Reality
(London: Routledge and Kegan Paul, 1963), p. 121.
GEORGE F. R. ELLIS

COSMOLOGY AND VERIFIABILITY

Relativistic Cosmology aims to determine the structure of the Universe from


a fusion of the results of astronomical observations with knowledge derived
from local physical experiments. The problem of determining this structure!
is centred on the fact that there is only one universe to be observed, and that
we effectively can only observe it from one space-time point. Because it is a
unique object, we cannot infer its probable nature by comparing it with
similar objects; and (on the scale we are considering) we are unable to choose
the time or position from which we view it. Our predicament is analogous to
that of a premaritime man living on a small island in an ocean, who observes
around him a host of other small islands apparently scattered at random on a
seemingly limitless sea. Unable to move from his island, his theory of the
world in which he lives can only be based on this partial view.
Given this situation, we are unable to obtain a model of the universe
without some specifically cosmological assumptions which are completely
unverifiable. Because we wish to talk about regions we cannot directly
influence or experiment on, our theory is at the mercy of the assumptions we
make. To illustrate this, consider the possibility that our friend on the island
might be of a theological disposition and might have decided that there was in
fact only one island in the world, surrounded by an ocean which ended at a
beautifully painted diorama constructed by an artistic and kindly God, giving
him the illusion of a limitless sea covered by islands. In an exactly analogous
way, a modem cosmologist who was also a theologian with strict fundamen-
talist views could construct a universe model which began 4004 years ago in
time and whose edge was at a distance of 4004 light years from the solar
system. A benevolent God could easily arrange the creation of this universe
not only so that suitable fossils would be present in the earth (having been
created, together with the rest of the universe, 4004 years ago) to imply a
long geological history, but also so that suitable radiation was travelling
towards us from the edge of the universe to give the illusion of a vastly older
and larger expanding universe. It would be impossible for any other scientist
to experimentally or observationally refute this world picture; all that he
could do would be to disagree with the author's cosmological premises.
What has been violated here is the expectation that the ordinary laws of

93
R. S. Cohen and M. W. Wartofsky (eds.), Physical Sciences and History of Physics, 93-113.
© 1984 by D. Reidel Publishing Company.
94 GEORGE F. R. ELLIS

everyday physics, carefully and correctly applied, will lead to correct


inferences about what exists; for such application of these laws leads (in these
unusual universes) to expectations different from what would actually be
there. To exclude this possibility, we invoke the first of our unverifiable
assumptions about the universe: whenever normal physical laws can be
applied, they correctly predict the structure of the universe. I shall call this
the local predictability assumption. There are two facets to this require-
ment: firstly, that the normal physical laws we determine in our space-time
vicinity are applicable at all other space-time points. This demand for
uniformity in nature is necessary for reasonable predictions to be made about
distant parts of the universe; for otherwise there is too much arbitrariness in
what we can suppose. Without this guide, we have no suitable set of rules to
tell us what to expect. In any case we have a set of physical laws which are
locally valid, and the established scientific policy - based on the 'Occam's
razor' or 'minimal assumption' attitude - is to continue extrapolating,
applying these laws in larger domains and to more distant points, unless
something makes it clear that this is the wrong procedure. Note that this does
not exclude, for example, theories in which the gravitational constant varies;
all this amounts to (assuming we know the law which determines how this
'constant' varies) is that local physical laws are rather more complex than we
might originally have supposed. The second aspect of this statement is the
implication that we keep on applying these laws as long as this is possible;
and the resulting expectations are fulfilled. It is this aspect (expressed
mathematically in the requirement that space-time be inextendible 2 ) which
prevents our universe model beginning or ending at an edge such as the one
described above.
Having adopted this principle, one might hope it would not be necessary
to make further unverifiable assumptions in order to obtain a reasonably
unique cosmological model from our observations. However, the nature of the
observations we are able to make prevents fulf:tlment of this hope. Two facts
lead to this conclusion. First, we are unable to examine directly space-time
itself or the distribution of matter in it; rather we observe particular objects
- stars, galaxies, quasi-stellar objects, dust, and so on - in space-time, and
only when we have somehow determined their intrinsic properties can we
deduce their distribution and the properties of the intervening space-time. 3
Second, we simply do not have the astrophysical information needed to
determine their nature sufficiently accurately. This is partly because there is
a wide variation in the properties of individual objects in each class; partly
because we simply do not understand the nature of some of the classes of
COSMOLOGY AND VERIFIABILITY 95

objects we are observing; and to a very large extent because the light we
receive from the more distant objects was emitted a long time ago. Thus we
need to have a satisfactory theory of their time-evolution in order to deter-
mine their intrinsic properties at the time they are observed by us. We do not
have such a theory. So, for example, having obtained measurements of the
radio brightness of a radio source, we are unable to determine directly from
our measurements whether we are receiving radiation from a bright source
emitted a long time ago, or from a weaker source which is relatively nearby,
or from a weak source which is very far away but appears anomalously bright
because of the curvature of the intervening space-time. The situation is similar
to that of the isolated man on the island if he is able to measure accurately
the apparent sizes of the other islands, but does not know their intrinsic sizes.
Any particular island he sees might be a small one nearby, or a much larger
one a long way off. A new principle is needed to order the observations.
As presented thus far, the argument may sound rather weak; it may seem
that introduction of a new principle is a counsel of expediency rather than
necessity. Might it not be that given sufficient time for increased under-
standing of the astrophysics involved, the problem would eventually simply
go away; for then we would have sufficient information to use the observed
objects as 'standard candles' which could be reliably used to chart the
universe? This is most unlikely to be the case, not only because of the nature
of the difficulties encountered in astrophysics, but because of one funda-
mental aspect of our present knowledge of the universe which has not been
mentioned so far.
This crucial feature is that the universe appears to be isotropic about us
to an extraordinary accuracy. In particular the number counts of distant
radio sources show that their average distribution is the same in all directions;
the X-ray background radiation is isotropic to better than 5%; and the
microwave background radiation is isotropic to better than 0.2%.4 No matter
what direction we choose in order to obtain information about the large-scale
structure of the universe, we obtain the same answer as for any other
direction. Thus there seem (on a cosmological scale) to be no preferred
directions about us; we are unable to point iJ;l a certain direction and say 'the
centre of the universe lies over there'; in fact we are unable to say that any
direction is particularly different from any other.
To consider the consequences of this, suppose our astonished friend on
his island found that his observations lead to the same conclusion. He would
then be able to use this fact to construct for himself models of his world,
even though he did not know the distances of the islands he observed. He
96 GEORGE F. R. ELLIS

(0)
• • •
• • • •
• • • •
• • ~
p •
• • • •
• ~
Q
• •
• • • •
• • • •
• • • •
• • •

• ••• • • • • •
(b)
• • • • • • • • • • • • • ••
• • • •
•• • • • • • • • • • • •
•• • • • • •
• ••
• •
• • •
• • • • • •
• • • • ••
• • • •
• • • •
• • • • ~ • • • ••
• • p •
• • • ••
• • • • •
• • • ••
• •• ~ • • •
•• Q • • • • ••
• .0 • • • • •
•• • • • • • • • • ••
• •••• • • • • • • •• •

• • • • • ••••

Fig. 1(a). An arrangement of particles which is statistically isotropic about every


particle (P and Q are equivalent). (b) An arrangement of particles which is
statistically isotropic about P, but not about any other particle (P and Q are not
equivalent).
COSMOLOGY AND VERIFIABILITY 97

would, after a while, discover there were two possible situations. Either the
islands could be scattered uniformly over a uniform ocean in such a way
that all islands were roughly the same distance from the island nearest to
them, and so that the world looked very much the same to any observer on
any island (see Figure lea»~; or they could be distributed in some other way,
for example, with all the islands that looked smaller a much smaller distance
from their nearest neighbors than all the islands that looked larger (see Figure
l(b». The common feature of all these other ways of arranging the islands
would be that they were all centred on his own island; by measuring the
positions of all the islands in the sea one would with complete certainty
deduce that his own island was at the centre of the visible part of the world.
Although he himself would not be able to point out any direction as the
direction to the centre of the world, an intelligent observer on any of the
other islands he could see would indeed be able to do so; and all such
observers would point at his own island!
• The situation in relativistic cosmology is precisely similar. We can con-
struct all space-times which would give exactly isotropic observations about
one particular galaxy; and they are either exactly spatially homogeneous and
isotropic space-times, which are isotropic about every galaxy; in this case, all
galaxies are eqUivalent; or they are centred on that one galaxy. This galaxy is
then at the centre of the universe. s The actual universe, which is not exactly
isotropic about us, may then be expected to be very similar to one or another
of these idealized possibilities.
In ages gone by, the assumption that the earth was at the centre of the
universe was taken for granted. As we know, the pendulum has now swung
to the opposite extreme; this is a concept that is anathema to almost all
thinking men. This is partly because we now believe that our galaxy is no
different from millions of others; more fundamentally, it is due to the
Copernican-Marxian-Freudian revolution in our understanding of the nature
of man and his position in the universe. He has been dethroned from the
exalted position he was once considered to hold.
It would certainly be consistent with the present observations that we
were at the centre of the universe and that, for example, radio sources were
distributed spherically symmetrically about us in shells characterized by
increasing source density and brightness as their distance from us increased. 6
Although mathematical models for such ~arth-centred cosmologies have
occasionally been investigated, they have not been taken seriously; in fact the
most striking feature of the radio source counts is how this obvious possi-
bility has been completely discounted. The assumption of spatial homo-
98 GEORGE F. R. ELLIS

geneity has inevitably been made, and has led to the conclusion that the
population of radio sources evolves extremely rapidly.7 What has therefore
happened is that an unproven cosmological assumption has been completely
accepted and used to obtain rather unexpected information about astrophysi-
cal processes.
It seems likely that reasonable theories will continue to make this
assumption. One may adopt this view simply because our own galaxy seems a
rather undistinguished place to be .the centre of the universe, or because of
deeper philosophic reasons. In any case we shall accept the implied attitude
and tum to consider the different ways it can be formalized. Important
differences in our concept of the universe arise if we formalize it in different
ways.
The traditional way of codifying the view that we occupy an average,
rather than a highly special, position in the universe is to adopt the Cosmo-
logical Principle: 8 that is, the assumption that the universe is spatially homo-
geneous. This principle implies the existence of a cosmic time, and states that.
all measurable properties are the same at the same cosmic time. In particular
our observations of the isotropy of the universe would mean that all other
observers viewing the universe at the same time would find their observations
equally isotropic. Hence one obtains as idealized universe models the exactly
spatially homogeneous and isotropic (or Robertson-Walker) space-times. 9
These are supposed to represent the smoothed out structure of the universe;
a more realistic universe model is to be obtained by superimposing small
perturbations on this completely smooth substratum.
The Cosmological Principle is a positive statement with far-reaching
consequences. An alternative way of proceeding is to make a negative state-
ment. Thus we might make the assumption: we are not at the centre of the
universe. (I shall refer to this as the Copernican Principle.) As has been
indicated above, this principle together with the observed isotropy of the
universe about us again leads us to perturbed Robertson-Walker space-times as
models of the observed universe.
To illustrate the differences between these two approaches, consider once
again our marooned natural philosopher. Having formulated for himself a
'cosmological principle' - that every part of the world is identical to every
other part - he triumphantly announces his homogeneous and isotropic
world model: the world is a completely smooth ball. Not only are all points
equivalent to each other, but for every point, observations made in any
direction are equivalent to observations made in any other direction. His
ladyfriend - who has been around all the time, but engaged on other enter-
COSMOLOGY AND VERIFIABILITY 99

prises - now correctly but somewhat unkindly points out that the world
doesn't look very uniform to her. 'This necessitates him explaining that the
world-model wasn't meant to be an exact model of the world, but only an
approximate one showing its basic, overall structure; a more adequate model
would be obtained by thinking of a lot of islands scattered all over the
idealized smooth ball. The homogeneity is meant to be understood in some
unspecified statistical sense.
As his friend's reaction is not completely positive, he broods overnight and
the next day formulates his 'Copernican Principle' - that their own island is
not at the centre of the world. He then easily convinces her that this principle
- not being stated as an exact requirement of uniformity - is readily amen-
able to a statistical discussion; and that (because of the isotropy of the world
about their island) it leads to the conclusion that the world they see is
approximately a smooth ball with islands scattered over it in a uniform way.
He is delighted to find she accepts the principle as compelling, and the
resulting world-model as an obvious consequence. The new formulation has
the advantage that unlike the Cosmological Principle which only applies to
highly idealized models of the world, the Copernican Principle can be applied
to realistic world models; and so is a more satisfactory way of formalizing
the assumption.
Nevertheless in practice these principles may be interpreted so as to lead
to the same ideas about the observed universe. The problem lies elsewhere, as
our friend realizes with a sinking feeling when his companion asks him 'Gee,
does that mean there are islands just like ours in all the parts of the world we
can't see?'. This question puts him in a quandary. His cosmological principle
made a definite prediction about all the unobservable areas over his horizon,
namely that conditions there are the same as conditions near him. But he has
no observational information whatever about these regions, nor will he ever
obtain such information in the foreseeable future; so this conclusion is a
direct result of his completely unverifiable assumption about the world. If
he merely assumes the Copernican Principle, this orders his world the same
way in the observable region, because he knows that in this region the world
is nearly isotropic about him. But he does not have any such information
about the unobservable regions, and accordingly the Copernican Principle (as
formulated here) makes no particular prediction about these hidden regions.
Indeed, according to the available evidence they could be totally different
from the areas near him. Thus there could be many more islands, or many
fewer, or no islands, or perhaps a continent in some part or other; or perhaps
his whole concept of the world as a roughly uniform ball might be wrong,
100 GEORGE F. R. ELLIS

for while it might have that form near him, it could be, for example, that the
region he saw was just the top of a mountain based on some landform of
completely unknown shape.
The situation in cosmology is essentially the same. The Cosmological
Principle determines a complete universe model; the Copernican Principle
only a model of the observed part of the universe. The first model is satisfying
because it is complete, but unsatisfying because it makes predictions about
parts of the universe which are beyond observation; one has only one's faith in
the integrity of this principle to validate these predictions. The second model
is satisfying in that it only attempts to state conditions in the observable parts
of the universe, but \s therefore also unsatisfying, as there are further regions
of the universe which it does not attempt to describe. Attempts to resolve
this by setting up some intermediate principle seem unlikely to help. For
example, we could postulate a weak cosmological pn'nciple: that we are at a
typical position in the universe; but the effect is essentially the same as that
of adopting the strict form of this principle (which is the form actually used
in most writing on the subject). One could alternatively give different formu-
lations of the Copernican Principle, such as 'we do not occupy a privileged
position in the universe', obtaining essentially the same world-models as when
using the original form of this principle. (A problem arising here is that it is
not absolutely clear what 'privileged' should be understood to mean.) These
alternative formulations do not solve the essential dilemma.
In order to be more precise, I shall briefly sketch the universe models
obtained when these two principles are used. In doing so, I shall take General
Relativity with vanishing cosmological constant A as the theory correctly
describing the effect of gravity and determining the structure of space-time;
similar models would be obtained if A'*<>, and from closely related theories,
such as the Brans-Dicke theory. I shall also take the conventional interpre-
tation lO of the observations, rather than one of the more exotic alternatives ll
(which explain certain puzzling features at the expense of introducing various
new problems of interpretation). To give an idea of these universe models,
consider the picture of the (curved) space-time obtained when two space
dimensions are suppressed; the resulting diagram is a space-time diagram with
only one space dimension, representing the total history of the universe. The
curvature of space-time results in the space-sections being represented by
curved lines; at each point we may think of the local time direction as being
orthogonal to the space section at that point.
When the Cosmological Principle is applied, the exactly homogeneous and
isotropic world models resulting can, in general, be represented as in Figure 2.
COSMOLOGY AND VERIFIABILITY 101

effective particle our galaxy's


horizon world/line

surfaces
of
constant
time

our past
light cone

_ _-1-- plasma

Singularity
at '-0
(R,)

Fig. 2. Universe model based on the Cosmological Principle.

(there is an exceptional case which we shall return to later.) The surfaces of


constant time are surfaces of spatial homogeneity, so all physical quantities
are identical at each of the events on anyone of these surfaces. The complete
histories of galaxies are represented by world lines, showing the spatial
position of the galaxy at each time; the separation of these world lines at any
one time represents the distance between the galaxies at that time. The
present time is represented by the surface t=to. The present expansion of the
universe is represented by the way the world lines intersect later surfaces of
constant time at wider separations; conversely as one considers earlier times,
102 GEORGE F. R. ELLIS

the galaxies are closer together. Continuing back in time, the matter particles
are ever closer together, and consequently the density of the matter is ever
higher; and within a finite time in the past, the density and temperature of
the matter become infinite at the initial 'big bang'. It is convenient to choose
the time parameter so this occurs when t=0. Mathematically, a singularity
in our world model occurs here; physically, known local theory breaks down
here: we are unable to predict to earlier times. Thus our universe model is
finite in time: the matter and radiation, and space-time itself, do not exist at
earlier times, so this represents the beginning or origin of the universe model.
The second important feature of these universes is that as well as the
matter, the radiation in the universe is compressed in the past, and becomes
indefinitely hot at early times. This means there is a finite time td
(to> td > 0) when the radiation is sufficiently intense to ionise the matter,
and at earlier times (Le., for td > t > 0) the universe is filled with a plasma
which is opaque to electromagnetic radiation. The third important feature is
our past light cone, i.e., the history in our past of light which we see at this
instant. 12 This represents the part of space-time from which we can now
possibly receive directly signals in the form of electromagnetic, gravitational
or any other type of radiation. It bounds the part of space-time from which
we could have received any signal or other form of communication because
particles and signals are unable to travel faster than light, and the light cone
represents signals impinging on us at the speed of light. Thus we can only be
influenced by events lying in or on this past light cone.
We can immediately identify seven regions in this idealized universe model
which have essentially different observational status. Region R 1 is the part of
our past light cone since the recombination of the primeval plasma (Le., for
to > t > td). This is the set of events from which we may receive information
by means of electromagnetic waves, in particular by light or radio observa-
tions, and by any· other form of radiation (such as gravitational waves). It
is the maximal part of the universe we can actually hope to see. (Part of
this light cone is blocked off from our view by intervening matter; we can
only actually receive radiation from a particular event on our past light cone
if nothing opaque intervenes.) Region R z is the part of our past light cone
prior to recombination (Le., for td > t > 0). While we cannot receive informa-
tion from these events by electromagnetic radiation, because the plasma
absorbs or scatters photons passing through it, we can in principle 'see' these
events by using very sensitive gravitational wave and neutrino telescopes.
Thus we can in principle directly probe this region by observing radiation
other than electromagnetic radiation.
COSMOLOGY AND VERIFIABILITY 103

Region R3 (the interior of our past light cone since decoupling) is part of
space-time which we cannot observe by any form of radiation. However, we
have sufficient information available (from our direct observations on our
light cone, and from other kinds of information such as geological data) to
be able to have a general idea of what conditions are like here. For example,
we see the Andromeda galaxy at a certain time in its history; by determining
its velocity, we can with reasonable certainty plot its previous motion, that
is, determine its world line in R 3 • In principle the same applies to the region
R 4 , the interior of our past light cone prior to decoupling, but in practice we
are unable to form a very precise picture of what is happening here because of
the damping effect of the plasma: fairly arbitrary initial conditions lead to
much the same final state, so conversely observation of the fmal state gives
rather little information as to the initial conditions.
Regions Rs and R6 are parts of the universe with which we can have had
no causal communication. The difference between these two regions is that
the galaxies whose histories are represented by world lines in Rs are galaxies
we could possibly have observed by light or radio waves emitted at some stage
in their history; whereas the galaxies whose world lines lie in R6 are ones
from whom we could never have received such signals. Thus while we can
predict something about the matter in Rs by extrapolating from our observa-
tions of this matter at earlier stages of its history, we have virtually no
information about the matter in R6 at any stage of its history, and so are
quite simply unable to predict the state of this matter from any observational
information available to us. Some of this matter could in principle have been
observed by gravitational wave or neutrino telescopes; but even when such
observations are feasible in practice, we will most probably obtain very little
information about the distribution of the matter from these observations.
The rest of the matter in this region could not have been observed by us by
any form of radiation whatsoever; nevertheless we could in principle detect
that some of this matter exists because of the effects (such as that due to its
gravitational Coulomb field) it has on our past light cone. However, no way is
known to decode these effects to determine what distribution of matter is
causing them; thus we cannot decide if a particular distortion is due to a large
distant object or a smaller nearby object. (This is a difficult problem involving
the 'constraint' equations of General Relativity.) So we cannot determine the
detailed distribution of matter anywhere in R 6 • Finally R 7 represents a
different form of unpredictability; it denotes the singularity at the origin of
this universe model, where the ability of known physical laws to predict
breaks down. This breakdown arises not because of a lack of data, but
104 GEORGE F. R. ELLIS

because attempts ,to predict using the local predictability principle and
presently established local physical laws lead to a contradiction. The universe
model may therefore be thought of as beginning at this time; the picture we
obtain throws no further light on this origin.
Provided we make one further assumption, the Copernican principle leads
to a rather similar universe model. The extra assumption we have to make to
ensure our universe model is reasonable is the Causality Assumption: 13 it is
not possible for an observer to encounter himself. Obvious problems
concerning the nature of free will arise in a space-time in which an observer's
world line can twice approach the same space-time point, so that he (as a
young man) meets himself there (as an old man). The assumption that this
cannot happen has to be made explicitly in this case. 14 (It was automatically

our past
I ight cone,
R!
our
past

Singularities
R7

Fig. 3. Universe model based on the Copernican Principle.


COSMOLOGY AND VERIFIABILITY 105

fulfilled in the exa,ctly homogeneous and isotropic space-times.) Having made


this assumption, we obtain a universe model (see Figure 3) divided into
regions R 1 -R, with the same significance as in the previous- model The
regions Rl and R 3 , and the parts of Rs near R 1 , are very similar to the
corresponding regions in the previous case; the present picture is just a more
accurate (so 'bumpy') representation of these regions than the idealized
(smooth) one. Thus the past light cone Rl may develop caustics and self
intersections; after such caustics or self intersections it no longer bounds R3
but rather lies inside R 3 • IS Our information is not good enough to let us have
precise geometrical concepts of regions R2 (the part of our past light cone
prior to recombination) andR 4 (the part of our chronological past 16 prior to
recombination) on the basis of the available data (again, part of R2 may now
lie inside R4). However, we may be confident of certain features; particularly
that there are very high density and temperature regions in R 4 , and that there
exists at least one singularity R, here. 1? The nature of this singularity is not
known; in particular it is not clear whether a large portion of the matter
crossing our past light cone in the universe model should be regarded as
originating at the singularity or not. There are some hints that these regions
are rather unlike a Robertson-Walker universe model,18 so this picture of the
universe will (when we have sufficient data available to make definite state-
ments about this region) probably differ considerably, in regionsR2 andR 4 ,
from the first picture. Finally we simply do not have enough information
available, on the basis of present or possible future observational data, to say
anything much definite about the region R6 and the parts of R 5 distant from
our past light cone. Continuity suggests that conditions in the parts of space-
time outside our light cone but near it, will be similar to conditions inside it;
but does not justify our holding the same expectation for more distant
regions.
Comparing these two pictures, it is clear that in the second, far less
sweeping use is made of the unverifiable assumption (the Copernican
Principle) than in the first (where the Cosmological Principle is the unverifi-
able assumption). We need to make some unverifiable assumption to order
our ideas about regions Rl and R 3 , as discussed previously; but the
Cosmological Principle has also ordered the overall structure (but not the
details) not only in the infinitely distant parts of R 6 , but also in parts of Rs
infinitely far away from us in time. These extravagant predictions about
parts of space-time completely out of reach of any form of observation seem
unreasonable; the impossibility of obtaining any relevant observational data
makes adoption of this ordering principle seem suspect, and rather
106 GEORGE F. R. ELLIS

arbitrary.!9 In any case, as mentioned above, there are some indications that
the uniform models might be wrong in regions R2 and R4 ; it may be that we
will be able eventually to prove that the Cosmological principle is misleading
if applied to these regions. It therefore seems that adoption of the Copernican
Principle is the better procedure.
Suppose then that we adopt the Copernican Principle, and so obtain a
universe model whose principle features are as sketched in Figure 3. We know
of the existence of the regions Rs and R6 because of the local predictability
assumption; but considerable uncertainty enters into what we will ever be
able to say, with reasonable confidence, about these regions, and about the
early parts of R 4 • The further away in space or time an event is, the less we
can reasonably hope to predict about conditions there. 20 If some regions of
the universe model are effectively beyond observational and experimental
research, what scientific status should we assign to these regions? How much
significance can we assign to these regions in our universe model, in this
situation?
When it was realized that knowledge of the microscopic domain was
limited by a fundamental principle of impotence (the 'Uncertainty Principle),
physicists took this principle seriously and made it the basic feature of
quantum theory. Should cosmology perhaps, as suggested by W. H. McCrea,2!
similarly take seriously the fundamental limitations on what we can say
about the universe, and turn them into the basic feature of our cosmological
theory? It seems likely that this is what we shall, in the end, have to do: to
acknowledge our inability ever to determine many features of the universe,
and to incorporate this indeterminacy as a basic feature in our universe
models.
At present we have no detailed proposal to hand for the implementation
of this idea. However, what we can do is to go back to our classical picture
of the universe, and examine in more detail the sorts of uncertainty that
arise. This will then provide the starting point from which further develop-
ments can proceed.
Let us refer back, then, to Figure 3. The initial uncertainty sets in on our
light cone: the further down our light cone we observe, the less we can say
about the objects we are observing. This is partly because of interference by
intervening matter, but primarily because of the distance involved: the
object has a smaller, fainter, more-red-shifted image if it is further away. The
amount of information we can obtain from observations of any particular
object in a given time with a particular telescope is limited by optical and
quantum considerations; and the further down our light cone the observed
COSMOLOGY AND VERIFIABILITY 107

object is, the less we can find out about it in a given time. 22 Despite the gen-
eral information we may eventually get from gravitational wave or neutrino
observations, it seems that we will never get a detailed picture of R 2 •
Our knowledge of local physics enables us to extrapolate from these
observations into R 3 ; if we consider the regions nearer to us in space and
time, the data on Rl is better and we can extrapolate back with more cer-
tainty to determine the previous conditions which have led to what we
observe. Some uncertainty arises because of uncertainties in the initial data,
and some because of the statistical nature of prediction both in quantum
mechanics and in statistical physics. We are also able to extrapolate back in
R3 to regions near our galaxy's world line in the very distant past, by use of
geological and geophysical data which tells us about the very early history
of our galaxy. This kind of information will probably give us our best indica-
tions as to conditions in R 4 , where physical conditions can be very extreme
and difficult to understand. In particular reasoning based on the 'cognizability
principle' developed particularly by Dicke and Carter 23 ,24 (we observe the
universe; so conditions in our past must have been of such a nature as to have
allowed the development of intelligent life) provides an interesting way of
deducing limits on conditions in the past.
Prediction in Rs is more of a problem. In principle, we should be unable
to predict anything here; for this is our observational future. It has not yet
been accessible to observation and we do not have complete data available
to predict what will happen here; indeed some of this region lies in our own
chronological future. General Relativity allows the possibility of arbitrary
electromagnetic or gravitational shock waves impinging on us without any
warning being givenCby data on R 1, and completely nullifying any prediction
we might make. In practice, this has not yet happened; and we may regard
it unlikely that it will happen, primarily because of the plasma in R6 which
shields us from any primeval disturbances. If some laser-type wave were
emitted towards us from the initial singularity, the plasma would attenuate
it and protect us from it; at the very least the diffusive effect of the plasma
would give us some warning of the approaching threat, in the form of a highly
increased black-body radiation temperature or distortions from a black-body
spectrum. In any case the large red shifts involved decrease the intensity by
an enormous factor. We have seen all the matter in the region R s , and can
therefore estimate what its future behavior is likely to be, and could hope to
deduce if it was likely to send high-intensity signals towards us. Thus the
overall structure of the cosmological model is such that local prediction into
the future is possible; the data we have (on R d is in practice sufficient to
108 GEORGE F. R. ELLIS

predict into parts of Rs near R 1, on an astronomical scale, because little


unexpected data arrives here from elsewhere in the universe. 2s Thus we have
no more difficulty in predi~ting where the moon will be, in 5000 years time,
than we have in determining where it was 5000 years ago. The main effect
of predicting into the future rather than the past is that our uncertainties
are somewhat greater; for example men might have destroyed the moon in
5000 years time, and a complete prediction has to include estimates of the
probability of this sort of eventuality. (The larger the scale of object con-
sidered, the less of a problem such possible interference is likely to be.)
Having accepted this somewhat greater uncertainty, no major difference arises
in predicting on a cosmological scale into the part of Rs near R 1 rather than
the part of R3 near R 1, from data on R 1. As we have only the idea of con-
tinuity to help in p~edicting into R6 and the distant regions of R s , we know
very little about these regions.
Putting this together, we arrive at an idea of how certain our knowledge
of various regions can be eventually. (yVe are more certain if we have more
information available, or if the reliability of our information is improved.)
The sort of picture we obtain is shown schematically in Figure 4; this indi-
cates the kind of accuracy with which we can hope to determine the structure
of the universe from observations, without introducing further unverifiable
assumptions. The mere acquisition of more detailed data will not change this
overall picture, but rather will fill in some of the details; it will only be upset
by some completely unexpected observational result changing our overall
view either of the concept of space-time inherent in General Relativity, or
of the cosmological data on which our universe picture is based;26 or by the
satisfactory inclusion in our theory of a new and -compe'1ling principle (such
as Harrison's 'bootstrap' principle 27 ) which necessarily orders those parts of
the universe beyond our powers of observation_ Without some such complete
overthrow of our present conceptual scheme, the diagram we have obtained
depicts the main features of the observable universe and the accuracy with
which we can hope to determine its details; the aim of observational cos-
mology, in this context, is to obtain the information necessary to fill in the
details of this picture up to the maximum attainable accuracy. The diagram
therefore represents the goal of observational cosmology, rather than its
present status. 28 It indicates the extent of knowledge of the universe one
can aim to have supported by observational evidence; and is therefore the
proper setting in which to consider how empiricist a view one should take
in constructing a cosmological theory. The somewhat mundane astrophysical
considerations we have taken into account are, in my view, essential to a
COSMOLOGY AND VERIFIABILITY 109

our galaxy's
world line

effective
particle
horizon

cone

.00001%

Fig. 4. Prediction certainty in the universe (schematic).

proper discussion of the epistemology and ontology of cosmology; I do not


claim to give such a discussion here, but rather to set the stage for it.
So far we have considered the general situation. Finally we should consider
the exceptional situation which arises if the universe has finite spatial sections
110 GEORGE F. R. ELLIS

(this necessarily occurs in the Robertson-Walker universe with k = +1, but


can also occur 29 if k = 0 or -1). The essential difference is that then there
are only a finite number of galaxies in the universe, and the universe has at
anyone time only a finite volume; and in many cases the universe has only
a finite future ahead of it. This alleviates the problem of discussing regions
R6 and Rs ; for there may not then be parts of the universe infinitely far
to the future of us, and there are no events indefinitely far away from us
spatially. In a restricted subclass of these idealized universes (in particular if
k = -1 and A very large) there may be no region R6 at all; our past light cone
may 'close up' on itself so that we could, in principle, obtain sufficient data
by optical and radio observations to predict completely the future and past
of the universe. In this case, no unexpected signals from unseen sources
would surprise us, as we would have sampled the history of all the matter
in the universe and no unexpected information can come from previously
hidden parts of the singularity R 7 • There would in practice still be consider-
able uncertainties as to what we could predict, but we would be able to relate
a major part of the universe to our observations in these cases; the observable
universe would be almost the whole universe.
How could we tell if this was the situation in the actual universe? The
crucial effect if the light cones closed up would be that we would be able
to see each galaxy in at least two different directions in the sky (unless some
opaque matter intervened). Thus in principle we have a simple way of seeing
if this is the case or not; we simply compare the images of galaxies in different
directions, and see if they might represent the same object or not. Unfor-
tunately this would be very difficult to establish in practice: not merely
because of the extreme difficulty of making the requisite observations (we
would expect this effect to occur at the limit of possible observations) but
also because the different images we would obtain of one galaxy would be
images resulting from light leaving the galaxy in different directions and at
different times in its history. Thus we would effectively be looking at the
galaxy from different directions and at different stages of its evolution. This
would make it virtually impossible to tell if one was in fact seeing the same
object in different directions in the sky, or not; nevertheless it is certainly
something one could attempt to do even if there is little hope that one would
obtain either a definite confirmation or a definite denial.
It is in fact more likely that the·question of whether or not these excep-
tional situations arise in realistic 30 universe models will eventually be decided
indirectly from evidence as to the amount of matter in the universe 31 and the
value of the cosmological constant, certain combinations of these values
COSMOLOGY AND VERIFIABILITY 111

making the exceptional situations inevitable, and other values making these
situations seem implausible (but not impossible). This is an important ques-
tion because of the major difference it makes to the verifiability status of
the universe; various authors (including Einstein and Wheeler) have argued on
ideological grounds that there must be finite space sections, and for a time it
was strongly argued, particularly by Sandage, that observational evidence
supported this view. The evidence from observations is not now widely re-
garded as being conclusive either way, and the question remains open; my
own somewhat biased view is that the present evidence makes it rather un-
likely that there are compact space sections, and very unlikely that the past
light cone closes up on itself. If this is correct, then the observable part of the
universe is a rather small part of the whole universe.
Perhaps the most intriguing question of all is the relation of cosmo genesis
- the nature of the singularity where at least part of the matter in the universe
is, in some sense, created32 - to observational tests. I suspect that definite views
on this will have to wait for a far more advanced theory combining a quan-
tum description of matter with gravitation, than any we have at present. If
some limits could be placed on the possible nature of the initial singularity
by some such theory, this might provide a further way of examining the
possible nature of those distant parts of the universe we have been consider-
ing. For in this case we would be beginning to understand the creation pro-
cess itself, and that ought to give us some ideas as to the limits of what might
be created. This or a 'bootstrap' argument 33 would enable us to progress
from merely observing the universe, to, in some sense, explaining it. At present
this is just a faint and distant hope - a gleam in the eye which may some
day come to fruition. But such a change in the technological or conceptual
apparatus available for examination of the problem could change the whole
situation and our whole certainty as to the nature of the universe. In fact our
friend on the island was last heard muttering profanely to himself as he
chopped down a tree and proceeded to fashion its trunk into a rudimentary
but serviceable boat.34

University of Cape Town

NOTES

1 An informative discussion of the theoretical basis of cosmology is given in H. Bondi,


Cosmology (Cambridge: Cambridge University Press, 1960). A useful discussion of the
history of the subject is in J. D. North, The Measure of the Universe (Oxford: Oxford
University Press, 1965).
112 GEORGE F. R. ELLIS

2 I shall, as discussed later, assume that General Relativity is the correct theory describ-
ing space-time and gravitation.
3 The way this can be done locally has been carefully described by Kristian and Sachs,
Astrophysical lournal143 (1966), p. 379.
4 A very readable discussion of recent observations and their interpretation is in D. W.
Sciama, Modern Cosmology (Cambridge: Cambridge University Press, 1971). A more
detailed discussion of the physics involved is in S. Weinberg, Gravitation and Cosmology
(New York: Wiley and Sons, 1972); a detailed review of cosmology and its philosophy is
E. R. Harrison's Cosmology (Cambridge: Cambridge University Press).
5 In certain such universes, there could be two centres. The argument then proceeds
unchanged; there are two galaxies whose situation is completely different from that of
all other galaxies in the universe.
6 Systematic redshifts could be observed in such an earth-centred universe even if it
were static; see Ellis et al., Mon. Not. Roy. Ast. Soc. 154 (1978), p. 187.
7 See Note 4 above.
8 The various principles are discussed in detail by Bondi and North (op. cit., Note 1).
Many other reviews of cosmology discuss them but in less detail, see e.g. the reviews in
Note 4; for a recent reappraisal, see Harrison's articles in Comments on Astrophysics
and Space Science 6 (1974), p. 29.
9 See the references in Notes 1 and 4, above, or the review articles in General Relativity
and Gravitation, ed. by R. K. Sachs (New York: Academic Press, 1971) (Proceedings of
the 47th International School of Physics "Enrico Fermi"); or in Cargese Lectures in
Physics, Vol. 6, ed. by E. Schatzmahn (New York: Gordon and Breach, 1973).
10 See Note 9 above.
11 See for example, the discussions in Nature 241 (1973), pp. 109 and 338-340; 242
(1973), p. 108.
12 On a cosmological scale, 1000 years is effectively an 'instant'.
13 More precisely, the 'strong causality assumption'. (See S. W. Hawking and G. F. R.
Ellis, The Large Scale Structure o/Space-Time (Cambridge: Cambridge University Press,
1973) for a discussion of these principles.
14 See K. GOdel's article in Einstein: Philosopher-Scientist, ed. by P. A. Schilpp (New
York: Harper Torchbooks, 1959) for some thoughts on causality violation and its
consequences.
15 There may also be parts of the boundary of R3 and R4 disjoint from the past light
coneR I andR2·
16 See The Large Scale Structure 0/ Space-Time (Note 13, above) for a detailed discus-
sion of causal concepts and the singularity theorems of Penrose and Hawking.
17 See Note 16 above.
18 See Note 9 above.
19 There is a fairly widespread tendency to adopt the view that assuming there is no
change in conditions (as in the Cosmological Principle) is effectively making no assump-
tion at all. This is clearly untrue.
20 See 0_ Heckmann and E. Schiicking in Onzieme Conseil de Physique Solvay: La
structure et l'evolution de l'univers, (Brussels: Editions Stoops, 1959), and F. Hoyle
in Rendiconti Scuola Enrico Fermi, XX corso (New York: Academic Press, 1960) for
early discussions of this feature.
COSMOLOGY AND VERIFIABILITY 113

21 W. H. McCrea, Nature 186 (1960), p. 1035; 187 (1960), p. 583;La nuova critica,
Cosmologia, llle serie (1960-1961).
22 See for example, A. W. K. Metzner and P. Morrison,Mon. Not. Roy. Ast. Soc. 119
(1959), p. 657; G. J. Whitrow and B. D. Yallop, Mon. Not. Roy. Ast. Soc. 127 (1964),
p. 315; D. H. Gudehus, Pub. Ast. Soc. Pacific 84 (1972), p. 818.
23 B. Carr and M. J. Rees, Nature 278 (1979), p. 605; B. Carter, in Confrontation of
Cosmological Theory with Observational Data, ed. M. Longair (Dordrecht: D. Reidel,
1974); R. H. Dicke, Nature 192 (1960), p. 440.
24 E. R. Harrison, see Note 8 above. The basic idea is to make predictions by consistent
application of the concept "nature is as it is because it is the only possible nature con-
sistent with itself".
2S Local physics is therefore a meaningful enterprise (for local physics relies completely
on the concept of an 'isolated system'; but we cannot isolate any system from gravita-
tional radiation).
26 The picture would have to be seriously revised if, for example, observations even-
tually prove the universe has a hierarchical structure, as has been suggested by de
Vaucouleurs (Science 167 (1970), p. 1203) and others; see, for example, Gold in Nature
242 (1973), p. 24.
27 See Note 24 above.
28 It is salutary to realize that different observers' present estimates of the value of the
Hubble constant still disagree to a marked extent about the probable range of values for
this number.
29 O. Heckmann and E. Schiicking in Gravitation, ed. by L. Witten (New York: Wiley,
1962); G. F. R. Ellis, J. Gen. ReL and Grav. 2 (1971), p. 7.
30 But idealized in being smoothed out; a really detailed universe model describing
details of possible black holes, worm-holes and so on may in general be expected to have
a non-trivial region R 6 •
31 See the discussions in the references in Note 4.
32 An interesting discussion has been given by N. R. Hanson, 'Some Philosophical
Aspects of Contemporary Cosmologies,' in Philosophy of Science, ed. by B. Baumrlm,
The Delaware Seminar, vol. 2, 1962-1963 (New York: Wiley/Interscience, 1963),
pp.465-482.
33 See Note 24 above.
34 Some of the ideas discussed in this paper have since been pursued further by the
author. In particular, the reader is referred to the following: 'Limits to Verification in
Cosmology', Annals of the New York Academy of Sciences 336 (1980), pp. 130-160;
'The Homogeneity of the Universe', Journal of General Relativity and Gravitation 11
(1979), pp. 281-289; and 'Cosmology: Observational Verification, Certainty, and
Uncertainty', South Africa'! Journal of Science 76 (1980), pp. 504-511.
DAVID HEMMENDINGER

GALILEO AND THE PHENOMENA:


ON MAKING THE EVIDENCE VISIBLE*

The history of the history of science reveals changing styles; for the late
nineteenth century Galileo was the model of the empirical and positivistic
scientist, formulating general laws a la Mach, as summaries of experimental
data. For more recent writers such as Koyre and Burtt, Galileo was a Platonist
whose revolutionary work sprang almost full-grown from his head and who
did not do the experiments described in his dialogues - fortunately, for they
would not have worked if he had done them. I do not propose to take up the
issue of the Platonism or non-Platonism of Galileo's mathematical science; the
issue here is that these writers argue that Galileo did thought-experiments in
the course of developing his theories, and referred to experience primarily as
a fmal check in order to be sure that he hadn't gone wildly astray. 1 This
rationalist picture of Galileo has been thoroughly criticized recently by Drake,
Settle, and others,2 who have shown that Galileo's notebooks give ample
evidence that he did do experiments at the time when he was probably
developing the analysis of freely falling bodies, and that these were reasona-
bly accurate. Furthermore, even some of the experiments described in his
dialogues give good results when done today. I shall argue for another, related
point, that in any event, Galileo's own writings, particularly the Dialogue
Concerning the Two Chief World Systems and the Two New Sciences, give
ample evidence that he considered experience essential for the foundation of
his science and not only as a check on the results.
One of the places where Galileo says that theoretical results should be
checked against experience is after the basic results of his treatise on uniform
acceleration are presented in the Third Day of the Two New Sciences (TNS). 3
I shall return to this later to argue that this must be taken in the context of
the entire dialogue discussion surrounding that treatise, and that one of
Galileo's main concerns there is to show how hypotheses are to be grounded
in experience. Conversely, he also argues in TNS and elsewhere that
experience must be interpreted, and that the hypotheses or theories which are
to be tested against experience are also the tdols by which we carry out this
interpretive activity. This may sound circular, but it is really not so, as I shall
show by looking at the way in which mathematical proofs are carried out. I
shall leave for another paper the equally important question of how the

115
R. S. Cohen and M. W. Wartofsky reds.), Physical Sciences and History ofPhysics, 115-143.
© 1984 by D. Reidel Publishing Company.
116 DAVID HEMMENDINGER

concept of experience which is involved in this process of interpretation


differs· from the everyday sense of experience. 4 The position for which I shall
be arguing is largely in accord with the conclusions of William Wallace in a
recent paper on Calileo, 5 in which he argues that Calileo's conception of
science is best described as a mixture of Aristotelian and Archimedean
elements. By this I mean that Calileo's idea of the proper way to begin a
science is essentially the same as Aristotle's, and that he demands of it the
same sort of demonstrative rigor as Aristotle sought. As Calileo himself
said, his innovation was to combine this with an Archimedean application
of mathematics to nature, and, specifically, to natural motion, which
Archimedes did not do. (This does not dismiss the question of Calileo's
Platonism, for we would have to consider the relation between Plato's and
Aristotle's concepts of science, but that is another matter.)
The question, then, is what is Calileo's conception of empirical evidence
and how did he think it was to be obtained? I shall argue that his two dia-
logues are in large part lessons on the right ways to analyze and interpret
experience so that certain principles of the science - definitions and explicit
assumptions - become evident. There are two main features of these lessons.
Part of the task is to reflect on ordinary sense experience to find out what it
really shows, as opposed to what it may seem to show at first. One example
of this is the study of various sorts of motion in the Dialogue Concerning the
Two Chief World Systems (Dialogue), 6 resulting in the conclusion that much
of what seems to be evidence for the motion of the earth turns out to be
evidence only for relative motion. The problem here is to decide how
experience is to be interpreted, and why one interpretation is to be preferred
to another. To answer this question we must tum to the other part of the
task, which Calileo did not spell out but which I believe he did recognize.
This is to understand just what the function of mathematics is in natural
science. We all know that Calileo said that the book of nature is written in
mathematical characters, but precisely what does this mean?7 I shall argue
that part of what it means is that the way to discover sound results in science,
that is, the way to obtain results which can be seen to be good explanations,
is to emulate the way in which mathematical results are obtained. Specifi-
cally, this means to use a kind of analysis which is similar in some respects to
the practice of mathematical analysis in the form which it took in ancient
Creek mathematics, with which Calileo was familiar. My argument here will
follow up some suggestions made recently by Hintikka and Remes in a book
on the meaning of ancient Creek mathematical analysis. 8 If I am right, then
one of the ways in which Calilean science mathematicizes nature is to make
GALILEO AND THE PHENOMENA 117

the activity of studying nature similar to the activity of giving mathematical


proofs.
At the risk of going over material already thoroughly covered by others, I
shall begin by speaking briefly about Galileo's treatment of sense experience.
There is the well-known passage in The Assayer where he makes what was
later called the distinction between primary and secondary qualities; he says
that there are some qualities without which he cannot even conceive an
object, while others seem to depend on the presence of the perceiver. 9 1bis
has led some philosophers, like Burtt, to say that Galileo makes many of the
qualities important to us in everyday experience to be "secondary, unreal,
ignoble, and regarded as dependent on the deceitfulness of the senses."lO This
is too strong; the theme which runs through Galileo's writings is that we have
to look twice, for what seems one way at first glance will often tum out to be
quite the opposite upon closer examination, but we need not reject parts of
our experience as being altogether deceitful. We also know that he frequently
appeals to sense experience as better than argument, claiming Aristotle's
agreement (Dialogue); he links it with necessary demonstration (Letter to the
Grand Duchess Christina); he makes analogies with terrestia1 sense experience
to explain the appearance of the moon and of sunspots (The Starry Messenger
and Letters on Sunspots)Y In none of these writings, moreover, does Gali1eo
argue that only the qualities later called the primary ones should be
important to us.
The preference for sense experience is stated several times in the Dialogue
both by the Aristotelian, Simplicio, and by Gali1eo's main spokesman,
Salviati. They say that Aristotle preferred sense experience to any argument,
and Salviati adds that however Aristotle presented his doctrine, he must have
obtained his results first by means of the senses (pp. 32,55,51). Salviati fully
supports this position, saying that "our discussions must relate to the sensible
world and not to the one on paper" (p. 113). Later, though, he seems to take
a contrary position when he praises those who, like Copernicus, "were able to
make reason so conquer sense that, in defiance of the latter, the former
became mistress of their belief' (p. 328). 1bis passage has led some commen-
tators to see Simplicio as the defender of the senses against Salviati's Galilean
rationality.12 Yet when we look at the context, the statement acquires quite
a different meaning, coming at the end of a dialogue in which, in good
Socratic fashion, Salviati leads Simplicio himself to work out the basic
features of the Copernican scheme. He does this largely by having Simplicio
draw diagrams which will account for some of the observed or reported
features of the planets, and having him obtain the heliocentric system as the
118 DA VID HEMMENDINGER

one most in accord with what is seen. The remark that the Copernican
account flies in the face of experience comes afterwards, when the third parti-
cipant, Sagredo, wonders why, if this is so simple, it wasn't accepted long ago.
Salviati points out that the real problem is the annual motion of the earth, for
it implies that there should be much more variation in the brightness of Venus
and in the apparent size of Mars than is observed, and there should also be
stellar parallax, though this is not detected. Salviati says that only because of
a better sense than natural and common sense has he been ready to accept the
Copernican system - referring at least in part to the telescope, which helps to
reveal why the brightness variations are not as much as initially expected (if
its evidence is to be trusted).
Before the discussion reaches that point, though, it is interrupted, and
in material added after the first edition was published, Galileo has Simplicio
return to the problem of diurnal motion. This digression does not further the
analysis of the Copernican system; it serves two other functions, I think.
One is to be a comment on methods of interpreting sense experience with the
aid of mathematics, and the other, to help cover up the fact that Galileo's
treatment of the problem of parallax, which follows, is not very precise or
well founded. The participants state two more objections to the diurnal
motion - that the turning of the earth should make the vertical become
inclined, and that this turning would make it impossible to see stars from the
bottom of a well, for they would whiz past the mouth of the well too
quickly. Simplicio himself answers the first objection, pointing out that it
would apply equally well to a ship circumnavigating the globe, though such
an effect is not seen. The second objection is more difficult for him, but he
quickly sees that the problem is just as great if the diurnal motion is
attributed to the celestial sphere. With a little help from Salviati he works out
a partially mathematical answer to the problem, learning that angular rather
than linear speed is what counts; this is perhaps his high point in the
Dialogue. This lesson was inserted here to underscore his earlier achievement
in beginning the construction of the Copernican system; the lesson referred to
diurnal rather than to annual motion because the latter is indeed a harder pro-
blem, and stellar parallax was not observed until the nineteenth century.
Some pages after this digression Salviati discusses the absence of visible
parallax, arguing that the stars are so distant as to make this parallax im-
perceptible. Once again, he makes quantitative estimates, but this time they
are far less precise. He estimates the angular size of stars of various
magnitudes of brightness, and assumes that they are the same size as the sun,
in order to get an estimate of the size of the stellar sphere, or of the distance
GALILEO AND THE PHENOMENA 119

of the stars. There is no check on this estimate of the size of the stars, of
course, but worse, Galileo has Salviati give a method for gauging the angular
size of stars which could not work at all. He says that he has held a thread
in such a way that it just obscured a star, and then by measuring the diameter
of the thread, he could get the angular dimensions of the star. He explains
carefully how to measure this diameter accurately, and how to correct for the
fmite size of the pupil of the eye, but this is an accuracy which does not deal
with the central problem, which is that the angular size of a star cannot be
measured in this way. I do not want to claim that Galileo was intentionally
misleading here, but only that he probably knew that this method could not
yield enough precision to be a good lesson in the use of mathematical
estimates in the analysis of experience. For that reason, he added the discus-
sion of diurnal motion, where the analysis succeeds.
During this discussion of the Copernican system, Galileo also has Salviati
suggest that the real motivation of Copernicus was his desire to present a
unified system of the heavens, rather than a number of separate accounts for
the various planets. This led him to see tme suppositions from which
appearances could be derived, rather than merely to seek hypotheses to save
the phenomena (p. 341).13 Copernicus and Aristarchus, then, deserve praise
for their bold leaps of reason, but Galileo claims to have experience to back
up his defense of heliocentricity, and this is better than reason alone. 14 The
question remains, how are we to find the tme suppositions, starting from
experience; how are we to come up with hypotheses which are not mere
hypotheses? That, according to Galileo, was the aim of Copernicus as well as
his own goal. How should we examine hypotheses in the light of experience?
Before showing what I think to be Galileo's way of resolving this pro-
blem, let us look at another example of Galileo's way of interpreting sense
experience. Early in the Dialogue (pp. 70ff.) Salviati claims that the moon
must have a rough surface, as Galileo reported in The Starry Messenger, for
otherwise it would not reflect light as it does. Simplicio insists that it must
be smooth and polished for it to shine as it does. Salviati uses an ordinary
mirror and then a spherical one to show Simplicio that the former looks dark
except from one direction, when illuminated by a distant source, and that the
latter also does not reflect light as the moon does (the claim that a mirror
could look dark when illuminated still surprises many students, I find). The
outcome of the series of demonstrations with the mirrors is first, that sense
experience refutes plausible argument and, second, that the moon must be a
rough, diffuse reflector if the comparison with the mirror and the diffusely
reflecting white wall is appropriate. In other words, we understand the way
120 DAVID HEMMENDINGER

that the moon reflects light by finding a similar effect with terrestial materials
such as white walls and then conclude by analogy, where there are like
effects, there are like causes.
Such arguments by analogy are tricky, of course, particularly where there
is no independent check on the analogy - and here one of the main issues is
whether it is ever appropriate to compare terrestrial and celestial objects.
Galileo himself points this out in The Assayer, saying that nature can pro-
duce its effects in many ways; it is possible to produce a comet-like
appearance by reflecting light in an oily streak on the surface of a glass carafe,
yet we do not think that there is such a carafe in the sky. 1 5 He uses this
example as part of an argument that the nature of comets is uncertain, for the
same sort of appearance could arise in several ways or be explained by means
of several analogies. There are other times when he does claim to know how
an effect is produced, though, and when he does claim more than a hypo-
thetico-deductive basis for his conclusions. To see what distinguishes these
latter occasions from the others, I propose to examine two sections of his
dialogues, the beginning of the Third Day of TNS, and the criticism of
William Gilbert at the end of the Third Day of the Dialogue; in each of these
we find a remark to the effect that "here are demonstrations which rank very
little lower than mathematical proof'.
To begin, let us tum to the concept of analysis, since I am claiming that
Gallleo's way of interpreting experience resembles Greek mathematical
analysis in some respects. Many writers have referred to Galileo's use of the
method of resolution and composition, or analysis and synthesis (such as
Cassirer, Burtt, Mctighe, Randall, Wiener 16 ). The texts to which they usually
refer are Dialogue, p. 51, where Salviati describes Aristotle's own method of
discovery as the "metodo riso[utivo", and the beginning of the Third Day of
Two New Sciences, the treatise on motion. There has been a long dispute over
the question of whether Galileo was simply using familiar labels for two
directions of reasoning, from effects to causes and from causes to effects, or
whether any of the specific content of Greek mathematical analysis actually
contributed to his work. Reference to these two patterns of reasoning was
very common during the Renaissance and afterwards, as Neal Gilbert has
shown, 1 7 and was sometimes little more than lip service. Part of the problem
has been to discover just what analysis meant to the Greek mathematicians
and to find out why it might work, and here I think that the recent work by
Hintikka and Remes, The Method of Analysis, helps.Is
These authors distinguish between what they call analysis in the sense of
the analysis of propositions, and in the sense of the analysis of configurations.
GALILEO AND THE PHENOMENA 121

They argue that the former is what is usually meant when one speaks of
analysis as 'working backwards', assuming that the solution has been found,
and asking, from what could this proposition have come. If A is the proposi-
tion to be proved, one assumes A and asks what implies A, finding, say, B (or
one asks what A imples, finding B, in which case there is the additional prob-
lem of the convertibility of the implication). The questioning continues
until one reaches something, X, which is already known, and which can serve
as the premise of a chain of inferences leading back to A, which is thereby
proved. However, Hintikka and Remes argue that, first of all, this account is
not adequately supported by a careful reading of the Pappus text which is
the locus classicus. Their claim has been disputed,19 but their other argument
is independent of it, and is the more significant one. First of all, it isn't clear
why 'working backwards' should be any easier than working forwards as in an
ordinary deductive argument, and in the second place, if this process of
working backwords is a matter of finding premises which imply the desired
conclusion, it isn't clear why a synthesis is needed, as Pappus says is required.
On the other hand, if it is a matter of finding propositions which are implied
by the desired conclusion, then why does one have the right to expect that
these implications are convertible, as they would have to be to yield a
synthetic proof? Finally, this entire treatment of analysis in terms of proposi-
tions ignores the specific features of geometrical reasoning. Typically, a
geometrical proposition is in the form: if something is an A, then it is a B, or:
if something is an A, then something else can be found which isa B (the
former are theorems; the latter, problems of construction). That is, they have
the form, (x)(y) . . .(z)[A(x,y, . .. , z) ='>B(x,y, ..., z)], or the form, (x)(y) . ..
(z)[A(x,Y, . .., z) ='> (Eu) (Ev) . . .(Ew)B(x, y, ... , z, u, v, ... , w)]. A geome-
trical proof typically begins by supposing A(a, b, . .. , c), with the variables
instantiated, and then goes on to deduce B(a, b, .. .,c), with the introduction
of new elements, if necessary. This is based on the techniques of natural
deduction, and what is important is that it is very often necessary to intro-
duce new individuals into the scheme even though they do not appear in
either antecedent or consequent of the conditional. In geometry, this means
making constructions, and we know that in most cases the hard part of the
proof is finding the appropriate constructions.
According to Hintikka and Remes, geometrical analysis is a specific way of
maximizing the chances of success in the search for the constructions to be
introduced. In it, one assumes both the antecedent and the consequent of
the conditional which is to be proved. One then instantiates the variables,
representing this instantiation by a figure and then, having given oneself as
122 DAVID HEMMENDINGER

much information as possible with which to work, looks for consequences of


all that is now taken as 'given'; that is, one seeks new relationships and new
constructions. These constructions will generally be discovered as dependent
on both antecedent and consequent, but because of the convertibility of most
geometrical propositions, with luck one will reach a construction which can
be seen to be possible on the basis of the antecedent alone (together with
prior theorems) and which is the first step in the chain leading from antece-
dent to consequent. If this is found, then generally the proof will follow. The
point here is that geometrical analysis works because it is a way of giving
oneself the most information possible with which to work, and not because it
is reasoning which goes in one direction rather than another. Analysis as
analysis of configurations, then, is a useful method because it reflects fairly
closely some of the procedures of the working mathematician (see the
Appendix for a simple example).
I shall now look at the treatment of naturally accelerated motion in the
Two New Sciences (pp. 197-213) in the light of this 'configurational'sense
of analysis. A couple of introductory comments are in order. First, while
the account of analysis just given is an attempt to relate it to the practices of
mathematicians, we must note that the practices are rather more flexible.
In particular, a mathematician often looks for a proof by working from both
ends toward the middle; that is, he may ask both what follows from what is
given, and also what follows from or is connected with what is to be shown
in conjunction with what is given. In such a case he is carrying out parts of
the analysis and parts of the synthesis at the same time. The second point
is that we must keep in mind that this part of TNS consists of both a Latin
treatise by 'the Academician' (Galileo), which is read by Salviati, and a dia-
logue in Italian, involving the three participants, Salviati, Sagredo, and Sim-
plicio. The treatise is Galileo's scientific text itself, while the dialogue dis-
cusses problems encountered in the text and clarifies objections to it. I want
to look at the text as the synthetic presentation of the results, and at the
dialogue as the analysis which shows how one could arrive at the results of
the text. The dialogue contains the lessons in method; we do not have to take
it as an account of Galileo's own steps of reasoning as he originally carried
them out, but it does offer instruction in how scientific propositions can be
supported or made evident in experiences.
This section of the treatise begins with a definition of uniform accelera-
tion as that in which increase of speed is proportional to time; it asserts that
sense-experience shows that this "agrees with the essence of naturally accel-
erated motion" (p. 191), adducing simplicity as a further justification. This
GALlLEO AND THE PHENOMENA 123

simplicity is not trivial, but we know that Galileo goes to some effort later in
the text to argue against another equally simple deftnition; it is the sort of
justiftcation which can best be offered after the fact. Since Galileo says that
the deftnition is correct because it leads to demonstrated conclusions which
are in accord with sense experience, this sounds like hypothetico-deductive
reasoning. I shall show that Galileo did not mean this, but rather that
experience could lead to virtual certainty that the deftnition is correct. This
happens in the ensuing dialogue, in which the deftnition is examined, and to
the extent that the deftnition is justifted there, this is the Aristotelian element
in Galilean science; that is, Galileo's notion of how to begin a science is in
accord with Aristotle's account in the Posterior Analytics I. 2. I do not mean
to say that their concepts of science are altogether the same; as I indicated
above (Note 4), we would have to examine just what the. concept of
experience is with which Galileo works and just how he thinks that
experience is to be interpreted.
Sagredo's ftrst objection is that of course any deftnition may be made as
an hypothesis, but this one does not seem to agree with experience. It implies
that even heavy bodies begin falling slowly, while everyday experience shows
that they immediately reach a high speed. Salviati deals with this as he does
with objections to the motion of the earth, by arguing that reflective
experience actually gives the opposite conclusion. Here he introduces another
sort of everyday experience; he proposes to gauge the speed of a falling body
by the effect of its impact. Asserting that there is some sort of direct relation
between speed and force of impact, he argues that since even a heavy weight
has little effect in falling a short distance onto a stake, its speed in that case
must be small, too. Thus, as he says, the same experience, better considered,
yields the contrary of the supposed result. This example sets the pattern for
the rest of the discussion: given that we have experiences, and supposing the
deftnition to be correct, what further experiences can we ftnd or construct
which will enable us to justify the deftnition on the basis of experience. The
experiences described by Salviati are like the constructions in the geometry
problem; they make clear the connection between what is given (everyday
experience of heavy bodies falling) and what is to be shown (that the deftni-
tion is correct). As in geometrical analysis, the necessary constructions are
found by taking both the experiences we already know and the conclusion,
not actually established, and using the two together to pick out the additional
elements of experience which will establish the connection between everyday
experience and the definition - or at least show that the initial objection is
not well taken.
124 DA VID HEMMENDINGER

This argument is followed by another in which Salviati asserts that we can


easily "penetrate this truth by simple reasoning" (p. 200). This new argument
is not present because Galileo had doubts about the strength of the preceding
one, I think, but in order to give another argument about experience, and also
to introduce a new and subtle idea. Salviati suggests that a stone falling from
rest acquires speed in the same way as one thrown up loses it. If the falling
stone acquires great speed immediately, then one tossed up should lose
speed and stop equally abruptly. There is nothing intrinsically absurd about
this possibility, but our experience shows that it is not so. Here we have an
indirect argument, that reduces a kind of experience which is difficult to see
clearly to one which we see more easily in everyday experience - as it is
indeed a little easier to see things tossed up gradually slow down and start to
fall. This is not a matter of reason alone without reference to experience, but
rather of reasoning about symmetry applied to experience. There is another
part of this example which is more nearly a matter of reason alone, but it is
not simple. Salviati speaks of continuous acceleration from rest, which he
says is infinite slowness. According to Drake,20 this Galilean conception of
motion and change is essentially different from that of fourteenth-century
mathematical physiCS, for that was based on an inadequate theory of pro-
portion derived from the arithmetical books of Euclid, while Galileo had the
general theory of proportion of Euclid V available to him. 21 Whether this is
completely correct or not, it is clear that Galileo is introducing a difficult idea
here, that a body may accelerate t1:J.rough infinitely many degrees of speed in
a finite time, an idea that he uses later in his proofs.
This apparently simple argument of Salviati's, then, presents a novel and
difficult concept of motion through reference to everyday experience. What
had at first glance seemed to be an a priori argument in favor of the stated
definition turns out to be quite the contrary when better considered. The
argument starts by simply asserting what it purports to be establishing, the
initial slow speed of a falling body, because it is actually using experience
to analyze the concept of motion in a new way. It differs from the first
argument not by being based on reason alone, but by clarifying one
experience through reference to another where the two are connected by
general notions of symmetry and continuity. (It is true that Galileo is getting
away with a very hasty introduction of a difficult notion of continuity here.)
They are both equally empirical in being framed in the light of what we can
readily see - once we know what to look for.
Following this, Sagredo begins to speak about the cause of acceleration
and responds to some criticisms by Simplicio. It seems like a potentially
GALILEO AND THE PHENOMENA 125

fruitful topic, but is cut short by Salviati who says that they are to look at
the attributes of naturally accelerated motion and not to inquire into its
causes. This passage is often taken to be Galileo's rejection of the study of
substantial causes in favor of studying functional relationships. However,
in the context it also makes clear that what is still at issue is the proposed
definition, and it is to be tested by looking for its justification in experience.
In fact, Salviati seems to be suggesting that Sagredo's reflections on the causes
of acceleration are to be put aside, together with the many other speculations
on the subject, because this reflection is unregulated by anything resembling
the practice of analysis. 22
Sagredo then returns to the definition to restate it: uniform acceleration
is that in which speed is proportional to distance traveled. This is the error
made by Descartes and the younger Galileo, among others. Simplicio concurs,
saying that he can agree that a heavy body acquires speed in proportion to
the distance fallen. As Drake points out,23 Salviati responds to Simplicio's
statement, which refers to experience, and not to Sagredo's, which is about
the definition alone. He says that if a body falls this way, then we can
establish a one-to-one correlation between its speeds during its fall through
the first two lengths, and the speeds during fall through the first four lengths
of the distance fallen. Each of the latter will be double the corresponding
former speed. If we accept an extension of Theorem II from the section on
uniform motion, applying it to instantaneous velocities which are acquired
during fall, then it follows that the two times of fall will be equal. 24 This
means that a body falls two units and four units of space in the same time,
which is true only if motion is instantaneous, which we know from
experience it is not. Thus the argument is not a purely formal one, but shows
only that Sagredo's defmition, taken by itself, implies results which contra-
dict experience. If Galileo's own definition can be shown to agree with
experience, then we know that the two defmitions are different. Salviati's
argument here is like a proof by contradiction: assume Sagredo's definition
to be correct, construct the experience which it would describe and show
that this experience is inconsistent with what we see. In other words, he
shows that no construction, starting from experience, could justify the defmi-
tion. It is striking that here, where Galileo might have· tried to give a purely
formal argument that the two definitions are different, he did not do so.
Instead he showed that the same kind of reasoning which had been used to
establish connections between everyday experience and the one definition,
could also be used to establish the impossibility of those connections for
the other definition. 25
126 DAVID HEMMENDINGER

This is the end of the clarification. Starting from the definition as


supposed, Salviati has sought and found several sorts of experiences which
were, in effect, constructed to serve as bridges between our general back-
ground of natural experience and the definition. Each time an experience
seemed not to agree with the definition, or an alternative conception of
motion seemed possible, it was shown on the basis of these constructions not
to be the case. While Galileo has not shown that the definition is correct with
absolute certainty, he has used it to bring out features of experience more
clearly and to show how these do at least point to the definition, once they
have been interpreted and reinterpreted with the conjectured definition as a
guide. The more the definition and its immediate consequences are used to
indicate experiences that are apparently inconsistent with it, the more it is
supported by those very experiences; this is the first part of wh'lt I mean in
saying that Galileo is trying to make the principles of his scien·.:.e virtually
visible in the phenomena.
Having restated the definition, Salviati next states Galileo's Postulate,
that the speeds acquired by bodies falling down inclined planes depend only
on the heights of the planes (p. 205). It is followed by another discussion of
the same sort as that following the definition, in which it is examined and
justified. The issue is slightly different; as a postulate, it must be made so
evident that conclusions based on it will not be merely hypothetical. Initially
this postulate is not as clear as, for instance, the Euclidean postulate that a
line can be drawn between any two points in a plane, so a dialogue is needed
to clarify the postulate.
It begins with Sagredo saying that, imagining smooth planes, perfectly
round bodies, and no obstacles, he has no trouble seeing that the postulate
is correct; good sense [illume naturale] tells him so. Salviati wants more than
plausibility and offers an experiment which will provide assurance "little
short of equality with necessary demonstration" (pp. 205-206) - this is
one of the two passages where he makes such a claim. He then describes the
thought-experiment in which he imagines a swinging pendulum, the bob of
which always rises to the height from which it was released, except for a
slight loss due to friction. This remains true even if a nail interrupts the
motion of the pendulum cord so that the rising and descending arcs differ in
radius. If the nail is so low that the bob cannot rise to the same height, it
winds around the nail, showing that momentum or a tendency to continue to
move upwards remains. From symmetry we see that the momentum acquired
in fall is the same as that required to carry the bob up the same arc, and since
the momentum is the same in rising along different arcs, so it must be the
GALILEO AND THE PHENOMENA 127

same in descending them. (As before, Galileo seems to be using momentum or


impetus, not precisely defined, to gauge speed.) Sagredo accepts this as
conclusive proof, but now Salviati cautions him against accepting too much.
Acceleration along arcs is not the same as along their chords, so we cannot
conclude that an object on an inclined plane would behave in just the same
way. Furthermore, if the arcs were replaced by a pair of straight lines, their
chords, there would be a comer where they met and the falling object would
rebound. Still, he concludes that if this obstacle were removed "the mind
understands" that the impetus, and hence the speed, depend only on the
height of the planes.
This is strange; how far does this experiment go in giving more insight
than Sagredo's initial good sense? It ends with virtually the same statement,
that if we imagine impediments removed, we can simply see the truth of the
postulate. Why describe an experiment which does not fit the conditions of
the postulate exactly, particularly when he could easily have described two
inclined planes so connected that there was no sharp comer at their junction?
We know from Drake and MacLachlan that Galileo did other experiments
with such planes, made with a curve at the lower end, so that a ball rolling
down would be launched horizontally.26 How is Salviati's experiment better
than Sagredo's quick wit which Salviati praises elsewhere?
The answer, I think, lies in what one has to do to make such a principle
as evident as possible and visible in experience. First of all, some demon-
stration is needed, just as a proof is needed in mathematics to display the
connections which a qUick wit may grasp at once. Sagredo's insight depends
on his understanding of motion - exactly the subject under discussion, so
that his intuitive grasp of the matter may be obscuring what should be laid
bare. Second, a demonstration should show its result clearly. We can imagine
a demonstration which would more faithfully represent what is stated by the
postulate, such as two planes with curved ends, properly joined, but this
would be relatively complicated and would have essentially irrelevant features
to distract attention from the main point. More steps in the argument would
be required to show that these details were actually inessential - for instance,
the planes with curved tips would introduce a 'mixed motion', compounded
of rectilinear and curved elements. There could, of course, be a series of
demonstrations with planes of varying heights and the same amount of curva-
ture at the tips to show that the curvature was unimportant when the curved
part was small in relation to the linear sections. This would continue to com-
plicate the issue, though, and the pendulum demonstration, on the other
hand, is a 'natural' one, being simply devised. The one thing which it does
128 DAVID HEMMENDINGER

show clearly and forcefully is that for this case, at least, the height of ascent
is equal to the height of descent for different combinations of arcs. It is the
immediacy and directness of observation which matters here; we can see that
height and not inclination determines speed, even though we are seeing arcs
and not chords, and no argument is needed to get us to see what is happening.
The result is not more than what Sagredo grasped through the natural light,
but it is presented in a way that makes it more readily accessible to anyone.
This accessibility is what the demonstration has in common with mathe-
matical proofs. Proofs make their conclusions evident by providing a series
of steps, each of which is quite directly seen, which lead from what is known
to the new assertion. Before the proof is given, the relevant relationships
are buried, although they may be grasped through an intuitive leap; after-
wards they are readily available. We 'see' a mathematical proof, but Galileo
wants US actually to see the postulate made clear in the pendulum demon-
stration. I suggest, in particular, that we can again think of this experiment
as being like a geometrical figure. We construct the figure according to what
is given and add whatever additional elements appear to be needed, and then
point out the important relationships. If we were to start with a pair of planes
meeting at an angle, we would quickly see that this wouldn't work and that
we needed to replace the corner by the curved tips of the planes. As I said,
though, this would require an argument to show that the change didn't
really matter and that the 'figure' was really still the same. This would be like
a figure with too many elements in it, none of which helped to connect the
problem with what was already familiar or known. The pendulum is different,
first, because it is simpler. Second, if we take the unobstructed motion of the
pendulum as the initial figure, we have something familiar and important -
we know that in this case, the bob rises to about the same height as that from
which it was released. We can now add additional elements, the nail in its
various positions, and see that something remains unchanged. Thus, the
pendulum connects a familiar constancy to the new, sought-after relation-
ship, while the planes with curved tips being unfamiliar from the start could
not be that. Even though the pendulum does not model the conditions of the
postulate exactly, it gives a vivid demonstration of the independence of
speed from inclination.
Another clarification is necessary here. To be more precise, in giving a
geometrical proof of some general proposition, we first instantiate it, so as to
talk about a single triangle, rather than about triangles in general, even though
our interest is still in the latter. When we draw the figure, we are not
producing the mathematical triangle, but only a material illustration of it,
GALILEO AND THE PHENOMENA 129

to help us keep in mind what is important. We do not, of course, measure the


sides with a ruler to establish equality or inequality but, looking at the figure,
reason about the mathematical object, basing the proof on what it is to be
such a triangle and finding, for instance, that the sum of the interior angles is
equal to two right angles. We must make the same sort of distinction with
the physical experiment. The pendulum we see is like the figure and it
illustrates or stands for the ideal physical object - a pendulum free oflimita-
tions like friction, for instance. The ideal physical object does not differ
from the actual one as much as the geometrical object differs from the geo-
metrical figure, since even the ideal physical object is conceived as material,27
but there still is this distinction, although it often gets obscured.
To see that this distinction does exist, consider the two different functions
which a physical thing, such as an experimental set-up, can serve. It can be
taken for itself, just as it is with all its actual complexity; this happens when
it is being used to illustrate some result already known. On the other hand,
the same object or set-up can be used as something in which we can see an
ideal situation. Here it can serve as a means of making visible some of the
steps in a demonstration about such objects. We find the parallel in mathema-
tics. Having proved Euclid I. 32, I may take a triangular object, assert that its
angles add up to two right angles, and then measure them to verify that this
is approximately so. This does not offer support to Euclid I. 32 but merely
illustrates it. In the proof of that proposition, though, I use the actual triangle
to stand for the ideal one about which I reason, or in other words, I idealize
the actual object. Here, too, the mathematical figure is taken in two ways. If
taken just as itself, it may approximately illustrate a proposition, or be
approximately described by a theoretical result. If taken as a means for
getting at a mathematical proposition itself, it is idealized. In natural science,
a ball descending an inclined plane with slight but measurable friction may be
approximately described by a theoretical result which abstracts from friction,
and we may say that it is a more or less good illustration of that theoretical
result depending on the degree of approximation. If we are trying to prove a
theoretical proposition, though, we look for set-ups which permit the
transition to the ideal most readily, so as to help in the process of analysis
and demonstration.
The point here is that a good idealization, or a good physical set-up for
the purposes of getting at an idealization, need not be the best possible
approximation. Later in TNS, Salviati discusses the kind of abstraction
which happens in idealization, when he responds to criticisms of the analysis
of projectile motion because it ignores air resistance and the finite diameter
130 DAVID HEMMENDINGER

of the earth.28 He compares this to what Archimedes does in The Quadrature


of the Parabola (I. 6, 7), where he treats verticals as parallel. This is not simply
an approximation, Salviati says, but would be strictly true under the
idealizing assumption of infinite distance from the center of the earth. Note
here that Archimedes did not choose to treat verticals as sides of a triangle
with a vertex at the center of the earth, nor does Galileo suggest that he
should, although that would be a better approximation. It would suffer, of
course, from mathematical and conceptual complexity, and furthermore,
little would be gained over the procedure of first treating the convenient
idealization and then separately considering how good it is as an approxima-
tion.
The pendulum is thus not only a physical set-up which corresponds to the
actual geometrical figure, but it is also taken as ifideal, like the mathematical
figure itself. Salviati's argument is about the ideal case seen in the actual one.
Since the test of a good demonstration is whether it shows the sought-for
relationships clearly, the pendulum may do this better than would the pair of
inclined planes, even though the latter provide a closer approximation to the
conditions of the postulate. This is precisely the case. In addition to the
essentially extraneous elements in the pair of planes with curved ends,
friction would have a much greater effect with them than with the pendulum.
This would make the gap between the real and the ideal greater in the case of
the planes, and hence there would be more difficulty in seeing the ideal
situation. This is different from saying that the planes would be like an overly
complicated figure; they would also be like a poorly drawn and less useful
figure than the pendulum can be. However poor the arcs of the pendulum
are as approximations to straight lines, they present the ideal case clearly
because of the height-constancy they display. For all these reasons, then, the
pendulum serves to make something visible and evident and is thus more like
a necessary demonstration than the planes would be, despite the future leap
of insight required to transfer the result from arcs to lines.
This justification of the postulate gives us an experiment which functions
like a figure in a geometrical proof; the examination of the definition, earlier,
gave us experiments which function like constructions in the figure of a
proof. Both serve to connect familiar experiences with the scientific
statements in such a way that the statements are made evident in the
experiences. We have gotten away from analysis with the pendulum demon-
stration because it is given to us ready-made, with no account of how we
might obtain such a demonstration or see that it is indeed suitable. There is
another passage, this one in the Dialogue, which combines features of both
GALILEO AND THE PHENOMENA 131

and shows how we might go about finding such an experiment to serve in a


way similar to the figure in a proof. In a discussion of Gilbert's study of
magnetism, we find the other reference to an experiment described as ''very
little lower than mathematical proof' (p. 408). This discussion begins with
Salviati's praise of Gilbert's work, saying that Gilbert's method has "a certain
likeness to my own". The method he then describes in Gilbert's work is the
hypothetico-deductive one; for instance, the earth shows a number of effects
which, we think, belong only to magnets, and hence the earth is a magnet. In
other words, all the observed effects can be understood by taking this
hypothesis as true. Yet just after this praise Salviati also fmds fault with
Gilbert for being too little of a mathematician; more knowledge of geometry
"would have rendered him less rash about accepting as rigorous proofs those
reasons which he puts forward as verae causae [of what he observes]"
(p. 406). In the context of the dialogue this doesn't seem to have to do with
any failure of Gilbert to measure or quantify, for Salviati's subsequent
account of his own work has little of such quantification in it. Gilbert's fault,
apparently, was that his non-mathematical method did not connect his con-
clusions carefully enough with his observations; mathematics should provide
the model for proving thin{!fo. The point which Galileo is making here is the
one I brought up at the beginning, that mathematics is important in his
science because it provides both a model and also a tool for the discovery of
good demonstrations. In TNS, he has Sagredo and Simplicio agree that
mathematics and not logic can teach the mind how to discover sound reason-
ing (p. 175); here he shows us more of what he means. We must keep in mind
that he is speaking particularly of how to find experimental demonstrations
of scientific results, and not how to arrive at the propositions to be proved.
Mter his criticism of Gilbert, Salviati goes on to give what Sagredo found
lacking in Gilbert's account, an explanation of the effect of an iron cap or
armature on a lodestone. We need not regard Salviati's account as an accurate
description of Galileo's own search and discovery, but we can take it as
Galileo's own description of what he thinks is important about the process of
discovery so far as it contributes to demonstration. To use a current term,
it is Galileo's "rational reconstruction" of his investigation. In it, Salviati
says that the iron armature which caps the lodestone does not change the
attractive force of the stone, for it does not act at any greater distance than
without the cap. The change must be in the way the force has its effect; if
there is a new effect, there must be some new cause (this is the sort of remark
which is so general that, taken by itself, could not help much in practice). He
goes on to consider what change in the kind of contact there may be when
132 DAVID HEMMENDINGER

the armature is added, and finds this change in the greater density and
uniformity of iron as contrasted to the stone. This greater density and
uniformity can be seen in several ways - he refers to the fmeness of a steel
edge and to the visible impurities in the stone - and it makes possible a
higher degree of contact between iron object and iron cap than between the
object and stone alone. His assertion or hypothesis is thus that a larger number
of points of contact between iron and iron is the reason for the stronger attrac-
tion. After this analysis of internal structure leading to that hypothesis - an
analysis with which we may find fault because it does not also treat the effect
which the iron cap has on the shape of the magnetic field - he describes the
experiment. A needle, Salviati says, adheres more firmly to the armature by
its blunt eye than by its point. If the needle is attached to the armature by its
point, it adheres no more firmly than when attached to a bare stone, the
possible contact being the same in both cases. If the needle is attached to the
stone (with or without the armature?) and a nail to the needle, and the nail is
then pulled away, the needle will pull away with it if the point touches the
stone, but will remain on the stone if its eye is there. Thus, the strength of
attachment is less at the point, where the contact is also less, whether this is
between stone and iron or between iron and iron. These experiments with the
needle are what Sagredo says are nearly as certain as mathematics.
This experiment could be called a crucial experiment in the sense of one
which decides an issue clearly. It is not so much something on which an
entire argument or thesis depends as it is a demonstration of what is claimed
in it; it makes visible what is said there. It is also more than an illustration of
something already established; in its context it shows that greater degree of
r.ontact is connected with greater strength of attraction. Yet my point is not
that we have an example of hypothetico-deductive reasoning here, with an
implicit prediction from an hypothesis, which is then confirmed by test. The
experiment with the needle shows something only because of the preceding
argument, including the discussion of internal structure, however inadequate
that is. Without this, the experiment could suggest that contact is somehow
significant but not be able to separate it from other effects. The experiment
alone is something lipon which one might happen by chance while trying
out different effects of a magnet, but it would then simply indicate that
some sort of explanation is in order. It might be one of a number of such
effects which together would call for explanation, but these effects would
not themselves direct the way to that explanation. This, of course, is part
of what is meant by Popper and others when they deny that there is a logic of
discovery.
GALILEO AND THE PHENOMENA 133

Let us contrast this with Gilbert's account oflodestones with armatures. 29


He describes a number of experiments done with capped and uncapped
lodestones, finding that iron clings more firmly to a capped one, although an
uncapped one seems to work at a greater distance. He finds that a piece of
iron is no more strongly magnetized after contact with a capped than with an
uncapped one, and so on. Yet he does not organize this material into an
argument showing how or why the cap has its effect, and does not search for
experiments which have demonstrative force, or at least he does not argue as
if he made such a search. He presents a series of observations which would all
have to be taken into account by any adequate explanation of the effect of
the armature, but they do not seem to lead to or support any particular
explanation, and it is probably for this that Galileo has Salviati's criticism;
Gilbert does not show how to integrate reason and experience.
Galileo's experiment is different, though. The hypothesis has been sug-
gested already; the significant difference between lodestone and armature is
the difference in uniformity of matter, and therefore, in degree of contact
which they have with the suspended object. The test to be made is: is it true
that the greater degree of contact produces greater magnetic force? How,
then, does the needle experiment acquire its demonstrative power? I suggest
that it is by being set up by the following sort of reasoning: Let us suppose
that the contact at A is greater than the contact at B and (what is yet to be
shown) that the force at A is also greater than the force at B. (Here we have
the start of the analysis.) What consequences of these two suppositions can
we find? Clearly, one is that a greater weight could be supported atA than at
B; we could find a weight which could be supported at A but not at B. Or,
alternatively, we could pull first on the thing with contact A, then on that
with contact B and compare the pulls, though that may not be simple to do.
However, if we could apply the same pull to both A and B, then the one
which gave way first would be the weaker. To have these two contacts we
need three things ... a few more steps and we come up with the needle,
touching the lodestone at one end and some magnetic object at the other.
Given that the contact between needle and stone is greater when the eye is
attached than when the point is, which is a matter of geometry, we can
predict the result and see that the prediction is confirmed.
Here we have an analytic procedure; antecedent and consequent of the
conditional are both supposed and used to fmd additional constructions -
the details of the experiment. We then set up the experiment, knowing
already what the degrees of contact for each end of the needle are, and
proceed to demonstrate the conclusion. This demonstration has near-mathe-
134 DAVID HEMMENDINGER

matical rigor because of the analytical method used to set it up, in which the
important relationships are brought out. We should also note that the analysis
is involved in discovering the demonstration itself, not in coming up with the
hypothesis. So far as Galileo did not take into account the way a conical
armature could, in effect, focus the magnetic field more narrowly, his
explanation is inadequate. Nevertheless, it still shows that the degree of
contact is important, whatever else may be involved, and that this is at least
part of the reason for the effect of the armature. Making the evidence visible,
then, means finding experiences or constructing experiments which, together
with the accompanying account, let us see the hypothesis working as an
account of what happens. Such an experiment bridges the gap between
individual experiences and universal laws or rules, so far as that gap can be
bridged. These experimental demonstrations never fully achieve the status of
being part of a mathematical proof because the gap between actual and ideal
remains, but they come as close as possible, and far Closer than a hypothetico-
deductive justification comes. The latter says that the hypothesis could well
be the correct account; the former says that it is being shown in the experi-
ment itself, so far as that can be done.
The function of this sort of analysis, similar in pattern to mathematical
analysis, then, is to enable Galileo to connect scientific propositions with
appropriate experiences. If I am right, we should read these sections of his
dialogues as lessons in how to interpret nature so as to make our definitions,
principles, and suppositions evident - that is, visible in the experienced
phenomena. I do not mean to say that this is the sole function which experi-
ments have for Galileo, for they also serve to illustrate results which have
been found and which are considered sufficiently confirmed already, and
they are sometimes used to provide support for an assertion. One of the most
familiar experiments, the one conducted with inclined planes to show that
Galileo's law describing uniform acceleration is applicable to nature, is of this
latter sort (TNS, pp. 212-213). It is a final check on the mathematical result,
although it does not have demonstrative power, being just an illustration of
it. Still, this does not mean that Galileo Claims validity for his law only
because it is confirmed in this experiment. In the dialogue preceding the
proof of this law, which I called the analytical part of the book, he has tried
to show that the foundations of his science can be seen to be valid at every
step along the way by reference to experience. He never suggests that only
this final verification is important. It is at this point in the text that he has
Salviati say that Simplicio is right to ask for experiments to test Galileo's
result, because this is appropriate in the mathematical sciences of nature, in
GALILEO AND THE PHENOMENA 135

which the practitioners "confirm their principles with sensory experiences,


these being the foundations of all the resulting structure" (TNS, p. 212,
revised translation). Drake has argued that this translation is the liest one
because it expresses the ambiguity which is present in the original text -
either "principles" or 'sensory experiences" could be the antecedent of
"these".30 What I have said here supports this view, for it would be
appropriate for Galileo to say that both the principles and the experiences
are the foundations of the science, where the experiences are those which
make the principles evident.
I have been emphasizing the analytic pattern in the search for demonstra-
tions or justifications of scientific principles. I haven't spoken about how
Galileo thought we should obtain hypotheses to be tested, though we know
that his manuscripts contain accurate accounts of experiments, some of
which probably aided the process of discovery. We can find a few hints,
though this really amounts to talking about analysis in a looser sense. Salviati
says, "for a new effect we must seek a new cause", which amounts to saying
'find what you need to find', but we could imagine doing the needle experi-
ment before having obtained the supposed explanation. In this case we would
be taking a complex experience not yet understood, and seeking relationships
among the various parts of it. This would amount to working with the
observed effects, treating them as 'known' elements to be connected with
other, actually known thing.<! in order to reach an hypothesis about them.
We might observe the result of the needle experiment without really under-
standing it, but conclude that degree of contact was relevant. The visual
appearance of the lodestone, with its impurities, could suggest that the degree
of contact of the stone and of the iron cap could well be different, which
could then lead to the proposing of the hypothesis to be tested. Although
this is a vague sense of analysis, it is still analysis in the sense of analysis of
configurations; it involves seeking relations by carrying out experiments, con-
structing new ones, and looking for the connections between the known
and not fully known parts. I don't want to push this too far, though, because
I am not suggesting that there is something like a general method of dis-
covery which Galileo is using, for that would be to claim too much about the
process of analysis.
I also do not claim that Galileo was clearly aware of the distinction
between analysis of propositions and analysis of configurations, particularly
since he sometimes speaks of it in the former sense. I do claim that, being
familiar with the practice of mathematics, he was able to transfer some of
the same methods to the study of nature, and that his treatment of experi-
136 DAVID HEMMEN DINGER

ments as somewhat like geometrical figures and constructions is one of the


reasons for his success. 31 Furthermore, I am claiming that while Galileo's
dialogues certainly should not be taken as a record of his actual process of
discovery, they do represent what he thought important to say about his
results and the justification of them. I have argued that we have a number of
cases where the central issue was how to interpret sense-experience, and that
Galileo's general approach was to do this by looking to further experience to
clarify the first one, to refer to experience to clarify or test results of concep-
tual analysis, and finally, to seek experiences in which the principle at issue
could be made visibly evident, so far as possible. This last meant both doing
experiments from which the principle would follow almost as a mathematical
result follows from its demonstration, and possibly using such a demonstra-
tion for the reinterpretation of everyday experience. For Galileo, empirical
evidence in the strict sense is what one could obtain by seeing nature in the
light of a practice which is like that of mathematical discovery.
I think that we could also describe this aspect of Galileo's work as being
his search for what Newton later worked out in more detail; namely, how to
"deduce hypotheses from the phenomena". Newton said, in a well-known
passage, "As in Mathematicks, so in Natural Philosophy, the Investigation
of difficult Things by the Method of Analysis, ought ever to precede the
Method of Composition.,,32 While he goes on to speak about induction, if we
look at what he does, we see that induction is a rather small part of the whole
process; it plays a role in formulating the hypothesis but the important task
is to do experiments which will then make the hypothesis evident. These are
the crucial experiments - they are not the test by which the hypothesis stands
or falls but the demonstration that the hypothesis has its roots in the
phenomena and is not arbitrarily chosen.
·As I said at the beginning, if this is correct, then a further question must
be taken up - how does this procedure affect the concept of experience
itself? If everyday experience needs to be interpreted and reinterpreted,
where does this end? What is the relationship between everyday experience
and its interpretations? I have argued that to use as the tools for the process
of interpretation the hypotheses which are at issue, is not circular, any more
than analytical methods are in mathematics. This still leaves open the ques-
tion of whether any of the rules or canons for the interpretive procedure
can be spelled out. There is also another, related question: how far can this
procedure be carried out? As I have described it, Galileo's science is founded
on a very strict demand for evidence. Is it possible to continue to meet this
demand as scientific theories become more complex and more remote from
GALILEO AND THE PHENOMENA 137

everyday experience? Is it necessary to do so, or is that important only in


the very early stages of mathematical science?
I do not propose to try to answer these difficult questions now. However,
if it makes sense to talk of analysis in this way, and to talk about deducing
hypotheses or suppositions from the phenomena, then we make heuristics
an important part of the philosophy of science, as some others have also
argued recently. This would be the art of finding hypotheses which could
be made visible in experience, and of finding experiments which could justify
hypotheses, and would not be only a matter of psychology or sociology.
Lakatos has written on heuristics, but when he comes to speak about what
constitutes a good heuristic in science, a "progressive research programme",
he ends up saying that "the direction of science is determined primarily by
the human creative imagination and not by the universe of facts which
surrounds US.,,33 This seems to me to be giving up too much to the psycholo-
gists and sociologists, and it ignores the possibility that the direction of
science could be determined in part not by the universe of facts but by the
universe, which provides us with experiences to be analyzed and understood.
I suggest that we might think of Galileo and Newton's emphasis on obtaining
hypotheses from the phenomena and making them evident in experiences, as
a concern with standards for good models in science; a good model is one
which does make the hypotheses or principles evident. From this standpoint,
not all formally equivalent models would be equally good; som~ would
provide more insight than others, and would more readily lead to further,
unanticipated results. I am suggesting, in other words, that, as Mary Hesse
says, a good model is like a good metaphor,34 and that to seek to make the
evidence visible, as Galileo does, is to seek what is likely to be a good model.
That, however, is a subject for another paper.

Wright State University

APPENDIX
As an example of analysis, suppose that we are given the problem: to double
a given square. If we simply start from the given square to be doubled, it is
not easy to see what to do next. Let us suppose that we are also given the
double square which is in fact to be constructed; we can then ask what
relations can be found to hold between them.
We must choose how to place the two squares in relation to each other;
the particular way we pick may suggest useful auxiliary constructions. We
could, for example, construct them as:
138 DA VID HEMMENDINGER

or

Figs. 1 & 2.
The auxiliary constructions which these suggest, such as lines connecting
vertices of the squares, do not tum out to be very useful. We could also note
the symmetry of the square, though, and construct the figure as:

D
F G
B

o c
I H
Fig. 3.
This figure suggests that the diagonals of the two squares might coincide, and
it is not hard to prove that A and C lie on FR, and B and D on IG, so 'that
we can construct the common diagonals:
Fr-_ _ _-,.

I
Fig. 4.
We then observe that the double square is divided into four equal triangles by
the diagonals, while the unit square is divided into two equal triangles by one
diagonal. It follows that each of the four triangles making up the larger square
is equal to each of the two making up the smaller one. Since all of these
triangles are similar, being isosceles right triangles, they are therefore also
congruent. It follows that the corresponding sides are equal, or that AC = FG.
This is the end of the analysis, since it gives us the desired relation between
GALILEO AND THE PHENOMENA 139

the two squares; the side of the double square is the diagonal of the unit
square.
This result has been obtained under the assumption that the double square
has been given. It remains to be shown that the construction of the double
square can actually be carried out starting from the given unit square and the
relation just found. We must, in other words, carry out the synthesis. This can
be done by building up the same figure, using only what is given, together
with results already known. ThUS:
F G

I
Fig. 5.
Given the unit square, ABCD, construct the diagonals, which meet at E.
Extend AC, BD, and mark off EF = AB. Show that triangle ABE is similar to
triangle FGE and hence that AB is parallel to FG. Construct GH, HI, FI in
the same manner, and conclude that FGHI is a parallelogram; show that it is
equilateral and contains one right angle, and is hence a square (all this can be
done with propositions from Books I and VI of Euclid). Conclude that
FGHI is the double of ABCD by the argument already used, since EF has
been made equal toAB.
To obtain the proof, we see that the figure is used in one way for the
analysis and then reconstructed from what is given, for the synthesis. Having
given this proof, we may also see that other related constructions are possible,
such as the one used by Plato in the Meno.

NOTES

* An eru:lier version of this paper was presented to the Boston Colloquium for the
Philosophy of Science in January, 1977. I wish to thank the commentator, Professor
John V. Strong of Boston College, and Professors Robert S. Cohen and Marx Wartofsky
for· their criticisms. I am also grateful to Professors Stillman Drake, William Wallace, and
WInifred L. Wisan for their suggestions. The collection of essays edited by R. E. Butts
and J. C. Pitt, New Perspectives on Galileo, Western Ontario Series, Vol. 14 (Dordrecht,
Holland: D. Reidel, 1978), has several essays which are related to this paper, particularly
those by Wisan and McMullin, but it appeared too late for me to use it.
140 DAVID HEMMENDINGER

1 For instance, "For it is thought, pure unadulterated thought, and not experience or
sense-perception, as until then, that gives the basis for the 'new science' of Galileo
Galilei", Alexandre Koynl, Metaphysics and Measurement (London: Chapman and Hall,
1968), p. 13. See also Giorgio de Santillana, Reflections on Men and Ideas (Cambridge,
Mass.: MIT Press, 1968), p. 175, "Galileo uses facts only as a check, as a discriminator
between necessary and wishful arrangement."
2 Thomas Settle, 'An Experiment in the History of Science', Science 133 (1961),
pp. 19-23; 'Galileo's Use of Experiment as a Tool of Investigation', in E. McMullin
(ed.), Galileo, Man of Science (New York: Basic Books, 1967), pp. 315-337; Stillman
Drake, 'Galileo's Experimental Conftrmation of Horizontal Inertia. Unpublished Manu-
scripts', Isis 64 (1973), pp. 290-305.
3 Galileo, Two New Sciences, trans. S. Drake (Madison: University of Wisconsin Press,
1974), p. 212. This translation is generally more reliable than the earlier one by Crew
and de Salvio. Both give the pagination of the Edizione Nazionale of Galileo's works
(Volume 8), and I refer to this with the italicized page numbers.
4 I discuss this in 'Experience and Its Idealizations in Modern Science' (Proceedings of
the Ohio Philosophical Association, 1980, pp. 25-36), in which I take up the question
of how the idealizations of mathematical science are related to manifest experience; this
issue has also been discussed by Edmund Husserl in The Crisis of the European Sciences
and Transcendental Phenomenology, trans. by D. Carr (Evanston: Northwestern Univer-
sity Press, 1970), Part II.
S 'Galileo and Reasoning Ex Suppositione: The Methodology of the Two New Sciences',
in R. S. Cohen et al. (eds.), PSA 1974, Boston Studies in the Philosophy of Science,
Vol. 32 (Dordrecht, Holland: D. Reidel, 1976), pp. 79-104.
6 Galileo, Dialogue Concerning the Two Chief World Systems, trans. by S. Drake
(Berkeley: University of California Press, 1967), 2nd revised edition.
7 See The Assayer in The Controversy on the Comets of 1618, trans. by S. Drake and
O'Malley (Philadelphia: University of Pennsylvania Press, 1960), pp. 183-184, for the
complete text. Galileo's Platonism and his anti-Platonism stand out when he says that
"Philosophy is written in this grand book - I mean the universe .... It is written in the
language of mathematics .... " This passage (which is not complete in Discoveries and
Opinions of Galileo) contains two references to what is written by 'some man', that is, to
fictions, and it compares these to what can be read in the universe itself. The language
of mathematics is important because it is universal; anyone can learn to read it, and
anyone who does can read the true account of things. Without this we are reading
imitations, or are becoming imitators ourselves, merely following someone else's words.
Hence the Platonic theme: out of the cave (Galileo says, the dark labyrinth), away from
the imitations to the true things. Here too is the non-Platonic position - by doing this
we can come up with the true account of things themselves, whereas for Plato, the best
we can do in a written account, at least, is to come up with a better story, a copy of
reality, rather than a copy of a copy.
8 J. Hintikka and U. Remes, The Method of Analysis, Boston Studies in the Philosophy
of Science, Vol. 25 (Dordrecht, Holland: D. Reidel, 1974).
9 The Controversy on the Comets, p. 309; Discoveries and Opinions of Galileo, trans.
S. Drake (New York: Doubleday, 1957), p. 274.
10 E. Pi. Burtt, The Metaphysical Foundations of Modem Science (1931; reprinted
New York: Doubleday, 1955), p. 90.
11 The last three of these are in Discoveries and Opinions. (See Note 9.)
GALILEO AND THE PHENOMENA 141

12 Such as Dudley Shapere, in Galileo: A Philosophical Study (Chicago: University


of Chicago Press, 1974), p. 127, despite his disagreement with Koynl about Galileo's
rationalism.
13 See Wallace, 'Galileo and Reasoning Ex Suppositione'.
14 Some of this experience, of course, came from Galileo's use of the telescope to
observe the satellites of Jupiter. He did not make more of this here, perhaps because it
would have raised too many questions about how far the telescope could be trusted, as
Feyerabend has suggested in Against Method.
15 Discoveries and Opinions, pp. 260-261; Controversy on the Comets, pp. 246-247.
16 E. Cassirer, 'Galileo's Platonism', in A. Montagu (ed.), Studies in Honor of George
Sarton (New York: H. Schuman, 1944); Burtt, Metaphysical Foundations; T. McTighe,
'Galileo's Platonism: A Reconsideration', in Galileo, Man of Science; J. H. Randall, 'The
Development of Scientific Method in the School of Padua', Journal of the History of
Ideas 1 (1940); P. W1ener, 'The Tradition Behind Galileo's Methodology', Osiris 1 (1936).
17 Neal Gilbert, Renaissance Concepts of Method (New York: Columbia University
Press, 1960), Part II.
18 The precise significance of analysis in Greek mathematics has been discussed exten-
sively, as Chapter 1 of this book indicates. I shall avoid the controversies here in order to
focus on what I think is valuable in this book; namely, its study of why analysis may
have worked in practice.
19 See, for instance, Appendix I to The Method of Analysis, by Arpad K. Szabo.
20 S. Drake, 'Impetus Theory Reappraised', Journal of the History of Ideas 36 (1975),
pp.27-46.
21 'Impetus Theory Reappraised', pp. 41-42; also Drake's introduction to his trans-
lation of Two New Sciences. See also Galileo's dialogue, 'On Euclid's Definitions of
Ratios,' in S. Drake, Gali/eo at Work (Chicago: University of Chicago Press, 1978),
pp.422-436.
22 There are several places in Two New Sciences where exchanges take place between
Sagredo and Simplicio while Salviati remains momentarily silent (p. 175, pp. 201-202,
pp. 225-226, pp. 310-312). As in this one, the general issue seems to be the proper
function of mathematics in natural science. Salviati's response to these is to bring the
discussion back to the concrete problems of the interpretation of experience by means
of mathematical analysis. Can we not take these passages as part of Galileo's lesson in
the right way to combine mathematics and experience'? Galileo's dialogues have fre-
quently been called works of art, like Plato's; a study of the interactions among the
participants would show, I think, that Sagredo is not only a spokesman for Galileo's
'middle period', and Simplicio not only the dogmatic Aristotelian or naive philosopher.
Galileo uses all three to present an account of methodology, and Sagredo and Simplicio
frequently present the two sides of science, the interest in mathematics and in
experience, which are combined by Salviati, in the proper proportions.
23 'Free Fall and Uniform Acceleration', in Galileo Studies (Ann Arbor: University of
Michigan Press, 1970), pp. 214-239.
24 On the application of this proposition to non-uniform motion, see Drake, 'Velocity
and Eudoxian Proportion Theory', Physics 15 (1973), pp. 49-64. Drake seems to take
two different positions in this work and in the one just cited; in the earlier one he says
that Galileo is not applying Theorem II, but is stating a new proposition, to be proved by
what follows in the discussion. In the later one, he suggests that Galileo may be
extending Theorem II to instantaneous speeds. The point, though, is that one could
142 DAVID HEMMENDINGER

prove such a proposition, whether Galileo has really done so or not, and so while he may
be guilty of leaving out some steps, he has not committed a logical error here. In her
thorough study of this part of TNS, Winifred Wisan argues that although Galileo's state-
ment of Theorem II makes no reference to uniform motion, his proof, which he says is
carried out in the same manner as the proof of Theorem I (pp. 192-193), therefore is
valid only for uniform motion, as is true of Theorem I. See her 'The New Science of
Motion: A Study of Galileo's De motu locaU', Archive for History of Exact Sciences 13
(1974), p. 284, Note 13. However, the analogy between the proofs of the two theorems
is only that the same sort of use of Eudoxian proportion theory is involved in each,
and not that uniform motion is tacitly assumed.
25 Many commentators have followed Mach in arguing that Galileo did claim wrongly
that the definition of velocity as proportional to distance was logically impossible. They
point out that it is not, since this gives an easily integrated differential equation, which
can be solved to obtain the expression for position, s =A exp (kt). However, while a
body might in general move according to this law, it could not fall from rest in a finite
time without instantaneous motion, for at t = 0, this already gives non-zero values for
position and velocity. No shift of origin or change of scale will get rid of the problem;
the body has already covered a distance and acquired a speed at the very start. Thus, the
definition does have the fault pointed out by Galileo, at least at the moment in which
the fall starts from rest. I. B. Cohen has pointed out essentially the same thing, in
'Galileo's Rejection of the Possibility of Velocity Changing Uniformly with Respect to
Distance', Isis 47 (1956), pp. 231-235. However, he argues that Galileo got his result
by incorrectly applying the Merton Rule. To say this is, in effect, to say that when
Galileo set up the one-to-one correspondence between the instantaneous speeds for falls
through two units and through four units, as Drake describes it, he was partitioning the
intervals in a non-uniform way. This non-uniform partition leads to a faulty integration,
if put into modem terms. That begs the question, it seems to me, for it is a valid
objection only if we already know that the integration is with respect to time, and hence
that the intervals should be partitioned uniformly with respect to time. Since Galileo is
examining the proportionality of speed to distance, he is, in effect, partitioning the
intervals uniformly with respect to distance, and that is quite appropriate. Or, to put
it another way, even if we object that this procedure is incorrect because of the non-
uniformity of the partition with respect to time, what that means is that the two
definitions of uniform acceleration are indeed inconsistent with each other, even if each
is a possible kind of motion.
26 See the Drake paper cited in Note 2 and J. MacLachlan, 'A Test of an "Imaginary"
Experiment by Galileo', Isis 64 (1973), pp. 374-379.
27 Galileo makes this point at the beginning of Two New Sciences, p. 51, although he is
arguing there for the difference between physical and geometrical idealizations because
the former involve reference to matter.
28 Two New Sciences, pp. 274-275.
29 William Gilbert, De Magnete, trans. by P. Fleury Mottelay (1893; reprinted New
York: Dover Publications, 1958), Book II, Chs. 17-22.
30 In 'Galileo on Sense Experience and Foundations of Physics', (Isis 68 (1977),
Pf·108-110).
3 For an example of Galileo's own use of mathematical analysis, see WlSaD; 'The New
Science of Motion', pp. 249-258.
GALILEO AND THE PHENOMENA 143

32 Isaac Newton, Optic/a; (4th Edition, 1730; reprinted New York: Dover Publications,
1952), p. 404.
33 I. Lakatos, 'Falsification and the Methodology of Scientific Research Programmes',
In Lakatos, I. and A. Musgrave (eds.), Criticism and the Growth of Knowledge
(Cambridge: Cambridge University Press, 1970), p. 187.
34 Mary Hesse, Models and Analogies in Science (Notre Dame: University of Notre
Dame Press, 1966), pp. 157 -1 77.
GEN-ICHIRO NAGASAKA

QUANTUM THEORY OF MEASUREMENT:


A NON-QUANTUM MECHANICAL APPROACH

1. INTRODUCTION

In the formalism of quantum mechanics we always find two statements which


refer to the term 'measurement' and which are generally regarded as basic
postulates of quantum mechanics, namely:
(1) If a measurement of the observable is made upon a system which is one
of the eigenstates of the observable, then it will give as result the corres-
ponding eigenvalue, and the state of the system will remain unaltered by the
measurement; and
(2) If a measurement of an observable is performed on a system, the state
vector of the system will be transformed into one of the eigenstates of the
observable.
The latter is often referred to as "the projection postulate", and let us call
the former "the eigenvalue postulate".
As is well known, the problem of quantum mechanical measurement has
arisen from the dual nature in the changes of state vectors which represent
microscopic systems. Normally a state vector changes continuously and
causally according to the time-dependent Schrodinger equation, but if a
measurement is made on the system, a transition to one of the eigenstates of
the observable, namely, a reduction of the wave packet, will take place
according to the projection postulate. The latter process is discontinuous and
apparently acausal. If we want to avoid von Neumann's epistemological
dilemma, the most natural path to take is to prove that the reduction of the
wave packet does occur according to Schrodinger's equation when a system
interacts with the measuring apparatus, so that the 'cut' of von Neumann may
be defmitely placed between the apparatus and the observer, and conse-
quently to show that it is not necessary to invoke the consciousness of the
observer to explain the reduction of the wave packet. This is equivalent to
saying that the reduction of the wave packet is a consequence of a special
type of interaction, and that the projection postulate is a theorem and not
a postulate of quantum mechanics. Accordingly the problem of measurement
will be reduced to that of construction of an adequate model of the measure-
ment process to which quantum mechanics may be applied, or, briefly,
145
R. S. Cohen and M. W. Warto/sky (eds.), Physical Sciencesand History o/Physics, 145-158.
© 1984 by D. Reidel Publishing Company.
146 GEN-ICHIRO NAGASAKA

that of fOrming a quantum mechanical theory of the measurement process.


Though many attempts have been made to work out a satisfactory
quantum mechanical theory of measurement, the gist of those attempts may
boil down to the proposal of the following model: first, the apparatus is
regarded as a quantum mechanical system and the measurement is a process
of interaction between the apparatus and the system to be measured;
secondly, the combined system of the two subsystems will change according
to the time-dependent Schrodinger equation, and after a certain lapse of
time the combined system will be in a mixture of the eigenstates of the
observable. Now suppose that at the moment of measurement the system is in
one of the eigenstates, i.e., 1/1W' of the observable, and the state vector of the
apparatus is M. Then if these systems interact, the combined system will go
over to the state:

(1)

disregarding the degeneracy of the apparatus states, where U is a Ul\itary


operator representing the time development of the combined system. It is
clear that this is completely in conformity with the eigenvalue postulate.
Thus if the state of the system at the moment of measurement is a certain
linear combination of the eigenstates of the observable, as a consequence of
(1) and from the linearity of quantum dynamics, the combined system will
be in a state:

U(t)[aw 1/l w ®M] =~aw[1/Iw®Mw]· (2)

But it has been shown by Wigner [1], and later more generally by Araki and
Yanase [2], that the process (1) is not possible unless the observable to be
measured is commutable with all the operators which represent conserved
additive quantities of the sYstem. Apparently this is too strong a restriction
on the observable for the latter to be measurable. However, they find
consolation in the fact that if the apparatus were large enough, equation (1)
would hold approximately. But Wigner has also shown that state (2) is not a
mixture but pure unless all the 1/1 WS except one vanishes. So it seems that the
quantum mechanical theory of measurement has ended in failure, and that
the only alternative is either to accept von Neumann's dilemma or more
radically to modify quantum mechanics itself. If we were to accept von
Neumann's dilenupa, we would be led to the sort of solipsism pointed out by
Wigner [3] and eventually to the idea of the Universal Wave Function of
QUANTUM THEORY OF MEASUREMENT 147

Everett [4]. But "Entification begins at arm's length; the points of condensa-
tion in the primordial conceptual scheme are thin~ glimpsed, not glimpses."
[5] The concept 'consciousness' may be talked about based upon our concep-
tion of 'things' which lie at 'arm's length'. Dearly without knowledge of
thin~ which are definitely objective, certain subtle statements about micro-
systems would never be asserted. On the other hand, however, quantum
mechanics has proved to be a good theory in its numerous applications, for a
vast range of microscopic phenomena, and it seems absurd to suggest its
modification on account of its failure in the measuring process. Since its
successful applications are based upon experimental data which have been
obtained by measurement, it is a matter of sound common sense that there
must be some explanation of how it is ever possible to measure quantities in
microscopic systems. Many attempts have been made to establish links
between the microscopic world and the world of common sense along this
line. The problem was "reconsidered and reinvestigated, in the spirit of the
'philosophy' of Jordan and Ludwig, on the basis of an ergodic theorem" [6],
yet not quite successfully.

2. SEMANTIC ANALYSIS OF QUANTUM THEORY OF MEASUREMENT

Quantum mechanics is a theory of physics which gives accounts of


phenomena in the microscopic world and yet it has been constructed from
and supported by a vast amount of empirical evidence which belon~ to the
macroscopic world and of which classical physics has failed to provide an
account. The main source of the peculiarities of quantum mechanics is clearly
the anomalous character of the empirical evidence on which it is based. The
correspondence principle has proved to be an extremely fruitful guiding
principle, and yet it will not be able to furnish any theoretical grounds for the
validity of quantum mechanics. The theoretical terms which appear in its
basic postulates are entirely alien to the vocabulary of classical physics in
terms of which the empirical evidence is expressed. Consequently quantum
mechanics possesses essentially the character of a self-contained and consis-
tent theory which is totally independent of classical physics and thus ready to
be formalized as an axiomatic deductive system. On the other hand, however,
the data which would furnish the empirical basis of the theory are described
exclusively in terms of the language of, say, macroscopic deSCription, and for
the specifications for constructing and operating any experimental apparatus
we need no vocabulary of quantum mechanics in any essential way.
Therefore, while the Camapian notion of the theoretical language L T and the
148 GEN-ICHIRO NAGASAKA

observational language La may not be applied quite precisely in most scienti-


fic theories, quantum mechanics seems to be an exceptional instance in which
such a distinction may be quite adequately made. And further, since the em-
pirical evidence is entirely stated in the language of classical physics, it would
be most natural to conceive of this as being the language La of quantum
mechanics, and the rest of the extra-logical terms which appear in the forma-
lism of quantum mechanics may be regarded as forming LT. Clearly such a dis-
tinction forms the core of various versions of the Copenhagen interpretation.
There may be different ways of formalizing quantum mechanics, but
whatever the case may be, we always find among its basic postulates the
two postulates mentioned earlier, or their equivalents. Clearly this is a unique
situation for empirical science with respect to the basic postulates which
refer to measurement; measurements are a means to obtain data and it is in
its relations with the data thus obtained that the plausibility of a theory
should be judged. Now in quantum mechanics the term 'measurement'
appears in its alleged postulates, and since the term is not defined within the
theoretical system, it may be taken as a theoretical term mT, a primitive of
the system. Oearly those postulates are statements which describe theoretical
processes. Being a primitive of the system, mT is understood only through its
association with these postulates, and consequently any statement which
contains mT must be regarded as a theoretical statement which describes a
certain theoretical process. In the case of classical physics, no postuhlte
contains the term 'measurement' and in whatever expressions the term may
appear, it is always eliminable in favor of other theoretical terms. Now in
empirical science, a theoretical statement must be confirmed either directly
or indirectly by empirical evidence. In the quantum theory of measurement,
however, mT is interpreted as also representing the actual process of measure-
ment. If we adopt this interpretation, the statement involving mT mus·t con-
tain expressions which belong to La. Since the statement involves mT in an
essential way, it must belong to LT and thus it must receive, in turn, empiri-
cal confirmation which is achieved only by means of measurement and the
latter is denoted again by mT. Thus we face the situation of infinite re-
gression and the whole theory will never reach empirical data upon which it is
actually based; the theory will float in the air. This conclusion may be
avoided only if the theory is trivial, or in other words, if every theoretical
term is defined in La and consequently may be eliminated. In this case mT
will merely designate a certain type of phenomena which are described in
terms of La. The above is exactly the situation in which von Neumann was
led to his celebrated dilemma.
QUANTUM THEORY OF MEASUREMENT 149

The quantum mechanical theory of measurement may be regarded as an


attempt to show that with an additional requirement being placed on mT, the
projection postulate may be eliminated while the theory remains consistent.
Thus it is a problem of theoretical consistency rather than an analysis of
the process of actual measurement of a microscopic phenomenon. Neverthe-
less, the two postulates seem well founded. But one must realize that the
eigenvalue postulate aims at establishing a connection between the theoretical
statements and the empirical evidence, and therefore should be more
adequately referred to as a rule of correspondence. The projection postulate
was introduced because the other postulates of quantum mechanics seem to
provide no account of why a result corresponding to a particular eigenstate
should be obtained as a result of measurement. Clearly it is not a rule of
correspondence. Of course, as mentioned earlier, it could be a theorem which
describes a certain theoretical process which gives an account of the
analogous experimental data. But if this were the case, an adequate model of
the process must be provided so that certain rules of correspondence may be
established.
Though a physical theory should not contain the term 'measurement' as a
primitive of its LT, it is quite possible and often plausible to give an account
of the processes of measurement. Since a measurement is a physical process,
it is only necessary that the theory be comprehensive enough to include such
a process within its explanatory scope. But a measurement of theoretical
quantity needs special consideration. Suppose a theorem which contains both
theoretical terms T1, T2, ..• , Tn, and observation terms 0 1, O2 , . . . , On, is
derived from a set of theoretical laws in conjunction with the correspondence
rule:

(3)

or in short

F(T, 0)= 0 (3 a)

Then, if a set of values, 0', 0", ... should be given to 0, it is, at least in
principle, possible to ascribe the corresponding values T', Til, •.. to T. It is a
matter of the design of the experiment which will make such determinations
of values possible. But it must be remembered that such ascription of T'S
to T is by no means to be taken as the definition of T. Here we must note
that the correspondence rule plays an essential role in the process of ascribing
150 GEN-ICHIRO NAG AS AKA

T'S to T, or in the measurement. Not every theorem which may be derived


from the theory in conjunction with correspondence rules corresponds to an
experimental process that is realizable. Therefore it greatly depends upon
the ingenuity both of the experimentalists and the theorists to design an
experiment corresponding to an adequate theorem or to derive such a
theorem so that the desirea values of a theoretical quantity may be obtained.
This again suggests the essential role of the correspondence rules for the
interpretation of the theoretical statement.
In sum, the quantum mechanical theory of measurement does not provide
an adequate model of the measurement of microscopic systems. It associates
erroneously the actual process of measurement with the eigenvalue and pro-
jection postulates, which should be taken as theoretical statements repre-
senting characteristics of microscopic phenomena. And it is clear that for the
analysis of the measurement process, the formulation of the correspondence
rules of quantum mechanics is indispensable.

3. THE CORRESPONDENCE RULES

Despite the fact that the fundamental and essential distinction between
microscopic and macroscopic phenomena has generally been accepted since
the proposal of the Copenhagen interpretation, it is surprising that so little
attention has been paid to the problem of their interrelations, or to the
corresponding rules. Undoubtedly the so-called Born interpretation provides
such a rule. But clearly it alone is not sufficient, since the eigenvector is a
complex function and is not determined solely by the probability distribution.
Of course if there is an uncontrollable disturbance of the system due to the
measurement, as is often mentioned by the proponents of the Copenhagen
interpretation, it will not be possible to establish such rules. But if measure-
ments disturb the system, then the effect of the disturbance must already
be included in the obtained data. Since the theory is constructed on the basis
of the data, if a comprehensive theory is ever to be possible, the possible
existence of a disturbance will not affect the relation between the theory and
data. And if the disturbance should really be uncontrollable, it would be
more reasonable to imagine that the theory would not exist at all. Thus there
must be rules which establish definite relations between the data and the
quantum mechanical statements, at least sufficient for the purpose of various
theories of microscopic phenomena.
Here the last remark may need some comment. Quantum mechanics is not
a physical theory in the proper sense of the word. It is rather, say, a theory
QUANTUM THEORY OF MEASUREMENT 151

matrix which is endowed with a potentiality for generating theories. If taken


as such, quantum mechanics possesses no explanatory or predictive force.
Being a partial differential equation, Schrodinger's equation will have definite
solutions only when appropriate boundary conditions are supplied. Thus it is
an essential requirement, for quantum mechanics to be a theory, that it
should be furnished with proper boundary conditions, and in order that the
theory may be tested, those boundary conditions must correspond to the
whole experimental apparatus including the system to be measured and its
operations, so that we may ascribe the corresponding values to the boundary
conditions as well as to the solution of the equation. Furthermore since the
validity and plausibility of quantum mechanics have been derived from its
numerous successful applications, the correspondence rules must be found in
those implicit and explicit assumptions which have been invoked therein.
Clearly the Born interpretation is a correspondence rule, but it establishes
correspondence only between the state vector and the statistical distribution
of empirical data; in this case, quantum mechanics provides only statistical
information concerning the microscopic events. But statistical information
should be concerned with an ensemble of systems which are mutually inde-
pendent. It is not relevant here whether or not such constituent systems
should be complex and include a large number of independent variables or
subsystems. If the measuring apparatus records the presence of a number of
microscopic systems, those data, say, clicks of the scintillation counter, or
dots on the photographic plate, should be regarded as representing different
microscopic events which are.mutually independent. Though the information
provided by the experimental data is of a statistical nature, still, we must
note, each of these data must correspond exactly to an individual event to
which a certain class of definite properties is ascribed. It is not the case that
to every event in the macroscopic world there corresponds a certain micro-
scopic event or vice versa, but there are certain macroscopic data each of
which represents an event which belongs to the microscopic world. Let me
call this the requirement of correspondence and accordingly, due to this
requirement, a correspondence rule will be established, which I will refer to
as the rule of correspondence. It is because of this rule that we may be able to
describe certain microscopic phenomena accurately, though allowance must
be made for the limit of experimental precision. Thus we speak of, for
example, a trajectory of an electron with a certain momentum. Is this a
correct statement? From the quantum mechanical point of view it is an
approximation, but by no means erroneous. An approximate statement is also
a correct statement? From the quantum mechanical point of view it is an
152 GEN-ICHIRO NAGASAKA

approximation, but by no means erroneous. An approximate statement is also


a correct one if we make allowance for a certain latitude of inaccuracy. In
fact all experimental data will involve a certain amount of inaccuracy, for
their accuracy is dependent on the limit of precision of the instrument
employed. Is the situation here radically different from the one ordinarily
involved in any experimental statement? It may be argued that from a
quantum mechanical point of view a statement like this is only about the
expectation value of the electron, and not about its actual state. Indeed it is
well known that the expectation values of quantities relating to an electron
in a potential field satisfy the laws of classical mechanics. But I think a mis-
understanding of the situation is involved in the assertion. What we are
concerned with is not the relation between the two theories, namely, the
quantum and classical theories, but the relation between a statement about
experimental data and a corresponding theoretical statement. It is absurd to
speak of the average value of the position and momentum of a single electron
at an instant of observation. Indeed, we do measure individual systems and
events, and it is these data that constitute the body of statistical information
about the microscopic phenomena which may be employed for the statistical
confirmation of, and prediction by, the theory. If a body of data provides
statistical information on an ensemble of systems or events, then each of such
data much represent an individual system or event as long as we postulate
that the constituent systems of an ensemble are mutually independent. Now
if each of the individual data should represent a microscopic system, the
state of the system must be represented by its theoretical counterpart which
possesses those characteristics exhibited by the particular data. It is the rule
of correspondence that establishes a specific correspondence between a
certain class of statements in La, namely, data, and a corresponding class of
theoretical statements, namely, state vectors. However, such correspondence
may vary according to the type of data and a particular rule of correspon-
dence may be formulated for each of the different types of phenomena. But
all in all, the relation between the data and the corresponding quantum
mechanical statements is exactly the same as that in classical physics. The
data are necessarily approximative but nevertheless they are correct represen-
tations of the microscopic states. The only requirement for the adequacy of
such correspondence is that the inaccuracy involved in data should not signi-
ficantly affect the consequences of the theoretical derivation for the data, but
apparently this requirement is not unique to microscopic physics and it is
again up to the ingenuity of the experimentalists to design experiments which
satisfy the requirement.
QUANTUM THEORY OF MEASUREMENT 153

A highly typical version of the rule of correspondence may be found in


the case of so-called 'state preparation'. In order to perform an experiment,
clearly it is necessary that a state be prepared. In most cases a preparation of
state is made to provide a set of initial conditions of the Schrodinger equation.
A preparation of state is performed by the operation of a experimental
apparatus. But the construction and operation of such an apparatus are events
which belong entirely to the macroscopic world, and the speCifications
required for the construction and operation of the apparatus will be described
completely in terms of the language of classical physics, La. And yet, if it is
possible that the state be prepared, the prepared state should be able to be
described in the language of quantum mechanics. Thus, there must be a
one-to-one correspondence between the speCifications and the eigenvector
assigned to the state. Take, for example, a scattering experiment. The
quantum mechanical state of the incoming particles must be determined in
conformity with the state preparation operation of the apparatus, and in
most cases it is a mixture of states which are the degenerate states of a
momentum eigenstate of the eigenvalue corresponding to the value deter-
mined classically by the process of state preparation. Thus quantum
mechanically, state preparation means a reduction of the wave packet, and
since the former is a process which, as far as the system is concerned, is a
microscopic process, a reduction of the wave packet should be taken as a
quantum mechanical process and accordingly, if stripped of the mysterious
measurement, the projection postulate should be taken as representing the
process, and as a theorem rather than a postulate of quantum mechanics.
Oose1y related to these two rules is another important rule, namely the
rule of the coherent range. The systems of which a statistical ensemble is
composed are normally separated in space and time. To assert that these
systems are mutually independent is equivalent to postulating that the space-
time extension of state vectors has a maximum range. As is well known in
classical optics, interference of light occurs only for a coherent range of wave
trains. It must be noted here that the notion of a coherent wave train is not
required as a boundary condition for the electromagnetic wave, but rather has
the characteristic of a correspondence rule which establishes a relation
between theoretical statements and experimental conditions. In fact, in the
actual treatment of interference and diffraction, this restriction is totally
disregarded. Similarly the rule of coherent range should not be taken as a
theoretical statement, but as a restriction on the experimental interpretation
of theoretical statements which makes the latter correspond to the actual
experimental situations. If microscopic systems are mutually separated by a
154 GEN-ICHIRO NAGASAKA

space-time distance beyond the range of coherence, they are regarded as in-
dependent and do not form a combined system but rather a statistical ensem-
ble. In a scattering experiment, for instance, the beam of incoming particles
is represented by a single state vector, normally with a defInite value for its
momentum. This does not mean that the state vector represents an ensemble
of particles which constitute the beam.
Here a statistical interpretation of quantum mechanics may seem tempting,
which asserts in effect that a state vector does not represent the state of an
individual system, but the statistical characteristics of an ensemble of systems.
But it is easy to see that such an interpretation is unacceptable. If that should
be the case, any linear combination of two-spin state vectors which are parallel
to the fIxed direction would never yield a state vector which represents a spin
direction perpendicular to the original direction. Furthermore, as suggested
by Lande, the crystal diffraction of electrons may be conceived as the prob-
ability distribution of momentum transfer between individual electrons and
the crystal lattice. But it is hard to see why we must take into account the
effect of the entire lattice to calculate the probability if individual electrons
are conceived as completely independent. Thus a statistical interpretation like
the above is not tenable, and the state vector must be taken as representing
the state of an individual system with the restriction of a limited space-time
range for its extension.
The correspondence rules are neither theorems nor axioms of a theory, but
rather postulates by means of which the theory is to be constructed from
the experimental data. The adequacy of the correspondence rules may be
judged by the coherence of all theoretical constructions and the scope of
application of the theory constructed. If the above statement is correct, then
the problem of measurement of the quantum mechanical system will be reduced
to that of the coherence of the theory and the correspondence rules. First,
since the data statements belong to Lo, and the experiment is designed and
conducted in accordance with classical physics, the behavior of the micro-
scopic system must be more or less in conformity with the laws of classical
physics of quantum theory. And it must be shown that there are cases such
that, when interpreted by means of the correspondence rules, the quantum
mechanical account of these events coincides with the account in terms of
classical physics, within the limit of experimental precision.
The allegedly anomalous behavior of the microscopic system could be
detected and talked about only on the strength of such coincidence for
those anomalies were expressed in Lo, and one must recall that it was due
to their existence that the need for a radically new theory was felt. Such a
QUANTUM THEORY OF MEASUREMENT 155

situation was at least in part the reason why the Correspondence Principle
was the only effective guide in the search for such a theory.
The assertion of such coincidence constitutes a most important part of
the rule of correspondence. It is rather strange that so little has been said
about state preparation. It is not measurement and yet we are sure that the
microsysterns prepared are in such and such states with respect to certain
observables. Without state preparation no experiment could be performed
which provides confirmation of a theoretical statement concerning a micro-
scopic system. Dearly the most important kind of measurement is founded
on the coincidence. The position of a particle can be measured and talked
about because it is there, and it is not brought about or created by measure-
ment. This means that the system in question is already in the state before
the measurement is performed. In comparison with the most sophisticated
arguments so far proposed this assertion may sound like a vulgarism, but I
believe that it is the only position the empiricist could ever defend.

4. THE POSSIBILITY OF CREATING A MIXTURE BEFORE


MEASUREMENT

If we adopt the terminology of the quantum theory of measurement, the


above assertion amounts to saying that the microscopic system is already in
the state of a mixture when it is to be measured. Now, since we are talking
about the behavoir of the microscopic system as such, the statement is a
theoretical statement in LT, or a theorem of the theory of quantum
mechanics. Dearly proving a theorem is a proper theoretical problem, and to
establish the consistence of quantum mechanics it is more than desirable to
show that the statement is a theorem of quantum mechanical theory. We
shall prove in the following that this is actually the case for a special model.
Suppose a microscopic system 91 coalesces and interacts with another
system 92 and is separated after interaction. Now let 91 be represented
by 1/1 and cf> respectively and the composite system S by '11m. We further
assume that '11m is an eigenvector of a Hermitian operator M in the direct
product space of the two systems, and thatM is conserved in the process. Then,

(4)
which, after interaction, goes over into

U(t) '11m = U(I/I ® cf»


= ~alJ.l/IlJ. ® 't(jvcf>p (5)
156 GEN-ICHIRO NAGASAKA

where 1/1/1, r/Jv are the base vectors of the respective Hilbert spaces and U(t)
is a unitary operation representing the time development of the composite
system, and

~ 1a/11 2 = ~ 1{jv 12 = 1
/1 v (6)
(1/1/1, 1/1/1') = D/1/1', (/1v, r/Jv') = Dw'
Since M is conserved,

[U(t), M] = o. (7)

Then the expectation value of M will be

(M) = ('It,M'It)
= (U(t) 'It, U(t)M'It)
(8)
= (U(t) 'It, M U(t) 'It)

= (~a/11/1/1
/1
® ~(jvr/Jv,M~,a/111/1/1' ® ~1{jVIr/JV')·
/1 /1 /1
Suppose that M is additive in the sense that

(9)

where Ml and M2 are Hermitian operators in the respective Hilbert spaces of


Sl and S2. If the 1/I/1's and r/Jvs are eigenvectors of Ml and M2 respectively,
then we may write

(10)

where

(10) and (11) show clearly that after interaction, Sl and S2 are in the state of
a mixture.
The above account, though it is based upon a particular model, indicates
that the formalism of quantum mechanics is in conformity with the sem-
QUANTUM THEOR Y OF MEASUREMENT 157

antics described in the previous section. The gist of the model consists in the
following three conditions; (1) the composite system before interaction is in
an eigenstate m of a certain observable M; (2)M is conserved throughout the
process; (3) the constituent systems are completely separated after interac-
tion. Though it is particular, the model is the commonest in application, and
therefore in practice it may be assumed quite generally that every isolated
system is represented by a mixture of states.
If the measurement is made on an isolated system with respect to certain
observables, and the rule of correspondence is applied to the data thus
obtained, we may say that quantum mechanics is a theory to give accounts of
the data, which is essentially a theory of the interaction involved. Since a
measurement always involves an interaction between the system concerned
and measuring apparatus, it is necessary that the interaction in measurement
should not be coherent with the interaction whose information we seek to
obtain. This is obviously realized if the two interactions are separated either
spatially or temporally by a distance beyond the coherent range. Further,
according to the rule of the coherent range, the measurement interaction
takes place only between the system to be measured and the microscopic
systems in the immediate neighborhood of the measuring apparatus, and the
neighboring systems interact with other neighboring systems, and so on.
. The latter processes constitute the amplifying process.
All these considerations show the vital importance of the correspondence
rules for the interpretation of quantum mechanics, and accordingly for the
analysis of the measurement of quantum mechanical phenomena. All in all,
the quantum mechanical theory of measurement is erroneous. Not only does
it stem from a mistaken interpretation of the projection postulate, neglecting
the essential roles played by the corresponding rules, but also it assumes
improper models for construction of the theory. If we consider the quantum
mechanical interaction between two systems, such a process in which the
state of one system remains unaltered while that of the other changes is very
unlikely or almost inconceivable. Hence the conclusion of Wigner and other
authors as to what should be expected.

Nanzan University, Nagoya, Japan

REFERENCES

[11 Wigner E. P. 1952. 'Die Messung quantenmechanischer Operatoren', Zeit. f Phys.


133101.
158 GEN-ICHIRO NAGASAKA

[2) Araki, H. A. and M. M. Yanase. 1960. 'Measurement of Quantum Mechanical


Operators.' Phys. Rev. 120 622.
[3) Wigner, E. P. 1967. Symmetries and Reflections, pp. 171 and 185. Bloomington
and London: Indiana Univ. Press.
[4) Everett, H., III. 1957. '''Relative State" Formulation of Qilantum Mechanics.'
Rev. Mod. Phys. 29454.
[5) Quine, W. V. 0.1960. Word and Object, p. 1. Cambridge, Mass.: M. I. T. Press.
[6) Danieri, A.,A. Loinger and G.M. Prosperi. 1962. 'Quantum Theory of Measurement
and Ergodicity Conditions.' Nuclear Physics 33 297.
JOACHIM PFARR

PROTOPHYSICS OF TIME AND


THE PRINCIPLE OF RELATIVITY*

INTRODUCTION

The discussion about 'protophysics' has hitherto been a purely Gennan affair,
restricted to the Gennan-speaking world. Nearly all publications, papers, pro-
grammatic declarations and also critical remarks have been written in
Gennan, and so American philosophers of science have been untouched by
this special version of a Philosophy of Science. Now since the book Proto-
physics of Time by Peter Janich (forthcoming)1 will be published in English
without a critical commentary, it is of some general importance to point out
publicly that there has been criticism of this program in Gennany and that it
comes from both philosophers of science and physicists. 2 The approaches of
the protophysical program have not gone unchallenged, as one can also learn
from the new edition or the English version of Die Protophysik der Zeit. 3
Because of the regional restriction of 'protophysics' it cannot be expected
that the English-speaking reader of this article is familiar with the proto-
physical program. Therefore before the details of this very program are
discussed, a short survey of the historical roots of protophysics, of its claims,
goals and intentions has to be given. And thus this paper is divided naturally
into three parts:
In the first part, I shall explain what has to be understood by 'proto-
physics' and a short historical survey is given.
The second part consists of a more detailed epitomization of what in
particular 'protophysics of time' is, and an immanent criticism of this parti-
cular program is included.
And in the third part an exemplary comparison between two different
approaches towards a definition of basic concepts and measuring devices in
physics is presented: the nonnative one of protophysics - based upon
instructions for actions - and the operational one of physics - based on
exemplary processes of nature.

I. THEPROTOPHYSOCALPROGRAM

The claims of protophysics are essentially based on the claims of physics


159
R. S. Cohen and M W. Wartofsky feds.), Physical Sciences and History of PhYSics, 159-189.
© 1984 by D. Reidel Publishing Company.
160 JOACHIM PFARR

to be a so-called 'exact natural science'. These claims of physics include:


(1) That physical theories - in contrast to, for instance, mathematical
theories - are per dejinitionem concerned with nature, with natural reality,
by no means with something merely produced in the mind or with self-
generated symbols.
(2) That physical theories have to be tested by experience, and are related
to experience.
(3) That the propositions of physical theories are 'exact' in the sense that
exactness not only refers to the accuracy of measurements possible at the
very historic moment due to the development of instruments and technology,
but also exactness in a more general sense. For instance, physicists will claim
that the propositions of Maxwell's theory are exact beyond the experimental
confirmation.
(4) That physical theories are objective. Here 'objective' means not only
objective in the sense of intersubjective testability; it also means objectivity
of physical theories beyond the linguistic, historical and sociological con-
tingencies of their genesis (ihrer Aufstellung).
As far as exactness and objectivity are concerned, these very far-reaching
claims of physics, considered in relation to the fact that its subject is related
to experience, give rise to the conjecture that parts of the physical theories
are based upon knowledge which does not originate from experience alone
but makes (scientific) experience possible and thus cannot be refuted by
means of experience. By this, however, are not meant the mathematical
theories ubiquitous in physics which help to describe nature and to formulate
theories more exactly. The knowledge in question explicitly refers to
experience without it being possible to derive such knowledge from
experience.
A model of knowledge about nature which contains these knowledge
claims as systematic components was first developed by Kant. The totality
of propositions concerning this knowledge - which he calls "synthetic judg-
ments a priori" - forms a physical frame theory which Kant calls
science" ("reine Naturwissenschaft") , "pure physics" ("reine Physik") or
"rational physics" ("rationale Physik,,).4 Some of these judgments a priori in
their role as a priori principles have been explicitly given by Kant, such as
the principle of causality and the prinCiple of the conservation of substance,
and Kant gave proofs for these principles in the Analogies of Experience. 5
It can be taken from the proofs of Kant for these two principles or from
other quotations, that the synthetical judgments a priori refer to a very
restricted domain of validity, a domain whose restriction primarily depends
PROTOPHYSICS OF TIME 161

upon Kant's interpretation of the concept 'experience'. For Kant, experience


is composed of something purely empirical and something a priori; experience
itself is not immediately given but is the unfolding of something given in the
senses by means of a priori forms of knowledge. These a priori forms of
knowledge - such as intuitions 0nschauungsformen), categories and trans-
cendental principles (Grundsiitze) - are a priori conditions of experience.
Given material of data can be woven into an experience only by means of
those a priori forms of knowledge. These forms playa constitutive or regula-
tive role when objects of experience are created from perceptions, and
because of this regulative role, experience tallies with the objects of
experience. 6 For the conditions of the possibility of experience then coincide
with the conditions of the possibility of the objects of experience, which
objects Kant calls "Dinge." 7 Kant's synthetical judgments a priori only refer
to objects of experience in this sense.
This is important to mention since some successors of Kant interpreted
the synthetical judgments a priori much less restrictively as presupposition-
less propositions about reality. The most well known of these philosophers is
Hugo Dingler who was concerned in his works with critical foundational
studies (kritische Begriindungsversuche) reaching from logic via mathematics
up to the exact empirical sciences, in particular, physics. The aim of Dingler
was to show "that it is possible to establish an enduring, so to speak, absolute
science apart from and logically independent of all sciences which have
become historical and conditioned by cultural contingencies."s As to
physics, Dingler hopes to answer the question how theories, i.e., systems of
propositions, can be applied to reality. Quoting Kant, he considers this as the
task of "submitting to principles (Grundsiitze) the transition from the
rational to the empirical.,,9 This aim he calls "full foundation" ("Vollbe-
griindetheit") and he tries to achieve this goal by means of what he calls the
"synthetic method," which method advances to secured results from a basis
by means of steps (e.g. logical rules) which by themselves can be traced back
to this basis. The principal problem then is what may be taken for the be-
ginning of this synthetic method as a basis and how to justify this.
Dingler takes as the basis a level which he calls "life standpoint" 1 0
("Lebensstandpunkt") or "standpoint of daily life" ("Standpunkt des tiigli-
chen Lebeni'). From this standpoint man has the talents for practical abilities
which make daily life, life as unreflected life, possible. According to Dingler,
these practical abilities are independent of any theoretical foundation.
Among all types of human actions, they include those which have to be
performed in order to construct an exact natural science, in particular
162 JOACHIM PFARR

linguistic actions and manual operations, which are necessary for the pro-
ductions of measuring devices. These 'basic abilities of man' include his
ability to make pre-scientific singular experiences and keep them. in memory.
Without going too much into details, we can say that this pre-scientific
knowledge along with the basic abilities of man to perform actions are the
basis of the Dinglerian system for the foundation of science.
According to Dingler, as a first step toward science, fundamental concepts
have to be defmed and measuring devices have to be built on the basis of
"instructions for action" ("Handlungsanweisungen")Y For instructions for
action, in contrast to pure asserting statements that need foundation, do
not need proofs. Instructions for action are on the same level as defmitions
and do not require any foundation. Nevertheless they contain statements
about reality.12
Using this methodological approach, Dingler tried to define the fundamen-
tal concepts of Euclidean geometry by means of instructions for action such
that the axioms of this geometry look like propositions which can be proved
by means of these operational definitions for the basic concepts. As an
example, only one of the operational defmitions for a basic concept of
geometry will be mentioned, that of the plane: Three disks A, B, C (roughly
smoothed beforehand) are rubbed against each other in the following way;
bottom of A against top of B, bottom of A against top of C, top of C on top
of B (C turned upside down before)P
Whereas Dingler's work contains many considerations as to the operational
foundation of geometry, the operational foundations of mass- and time-
measurement remain fragmentary. It is the merit of Paul Lorenzen to have
formalized and extended Dingler's considerations about geometry and mathe-
matics. More elaborately and more accurately, proceeding from logical and
linguistic considerations, Lorenzen in his investigations follows Dingler's
proposal and 'defmes' the basic forms of geometry by means of 'indistin-
guishability requirements" (or "homogeneity requirements,,).14 Thus a
sphere is characterized by means of the indistinguishability of all its points,
a plane by the additional indistinguishability of its two sides. Moreover,
Lorenzen formulates the program which he calls "protophysics":
Geometry, chronometry and hylometry are a priori theories which make empirical
measurements of space, time and materia possible. They have to be established before
physics - in the modern sense of an empirical science - with its hypothetical fields of
forces can begin. Therefore I should like to call these three disciplines by the common
name: 'protophysics'. The true sentences of protophysics are those sentences which are
defendable on the basis of logic, arithmetic and analysis, deimitions and the ideal norms
which make measurements possible.
PROTOPHYSICS OF TIME 163

We are dealing with materia, grinding its sides, regulating its movements and pro-
ducing collisions. We prescribe by norms how the materia shall "behave', if I may use
this biological metaphor .... [WJ e now force the materia to fit our ideal norms. In pro-
tophysics our relation to the world is no longer passive, we are now actively changing the
world. 1S
As to geometry in particular, Lorenzen says:
We define as geometry a system of theorems which can be derived from the norms which
deime the forms. Only their forms and not their realizations are objects of geometry.16

Thus protophysics, though - as Lorenzen claims - exclusively composed of


a priori theories is, insofar as the norms are concerned, based on the ex-
periences and abilities of man in the pre-scientific life world, where instruc-
tions for actions can be performed. Then the a priori character of the theories
can only be maintained if at least those parts of the life world, which have to
be used in order to perform instructions for actions, are of a priori character
themselves. Indeed the authors of the protophysics approach distinguish
between pre-6cientific experience in the life world on the one hand and
'empirically founded' scientific experience, based upon an experimentally
measuring physics on the other hand. (If in the follOwing their shared views
rather than the views of an individual author are relevant, they will be called
'protophysicists'.) By 'empirically founded' they understand those founded
assertion propositions (solche begrij,ndeten Behauptungssiitze) which can be
reduced to the measurement results of length, time and mass. And the a priori
then is interpreted as non-empirical in this sense. And they say that "it there-
fore makes sense to characterize the stock of experience in the life world as a
'life world a prion"'." 17 Such a less restrictive interpretation of the a priori
enlarges the sphere of a priori true propositions, which propositions now
not only include the analytically true propositions, the propositions of
mathematics, the judgments a priori formulated by Kant, but also "the
elementary propositions founded from the experience in the life world and
propositions logically composed thereof' ("die aus lebensweltlicher Erfahr-
ung begriindeten Elementarsiitze und /ogisch daraus zusammengesetzte
Siitze").1 IS
From this we learn that the claims of the protophysicists deviate from and
exceed the claims of Kant. Kant's synthetical judgments a priori refer only
to the restricted concept of experience as the unfolding of perceptions by
means of a priori forms of knowledge as intuitions, categories, principles. In
this sense they are valid universally and with necessity, but they are not
logically necessary and presuppositionless. In his 'pure science' Kant only
wanted to give the categorical frame of physics.
164 JOACHIM PFARR

Interpreting parts of the life world as a priori, the protophysicists more-


over claim to give a full foundation of all basic concepts of physics and to
proceed to presuppositionless propositions about reality. The whole of these
propositions then forms a new 'pure physics': protophysics. However,
because of these pretentious claims exceeding those of Kant, the suspicion
arises as to whether the ostensibly presuppositionless propositions about
reality might depend on conditions which are not discussed by the proto-
physicists in this context: the conditions of the possibility of performing the
protophysical instructions for action. 19
Whether this suspicion is justified or not can best be decided by carefully
investigating one particular part of the protophysical program and looking for
hidden presuppositions of the allegedly presuppositionless propositions of
protophysics. The most compact and complete part of the hitherto published
protophysical program is that concerning chronometry, done by Peter Janich
in his book Protophysics of Time. (See Note 1.)

II. PROTO PHYSICS OF TIME

Proceeding from a pre-scientific concept of motion and assuming that


Euclidean geometry is already completely founded (in the protophysical
sense), Janich develops a clock-free kinematics and furthermore a chro-
nometry. Clock·free means that he does not need clocks already defined by
special physical processes in order to derme kinematics. He defines clocks by
means of instructions for action. Deviating from Dingier, Janich replaces
periodic motion as basic motion for time measurement by straight uniform
motion. According to Janich, this is necessary because periodical motion
cannot be dermed without the concept of 'periodical motion' previously
being given (PoT, III, 4.1). By means of several definitions and propositions
about motion, avoiding any reference to concepts or devices which do not
originate from pre-scientific experience, Janich introduces straight uniform
motion by means of instructions for action. And finally he ends up saying
that his "methodological chronometry ... formulates requirements which
enable the realization of Galileo-invariant reference systems" (PoT, III, 4.41).
And Lorenzen agrees, saying:
If we inuoduce a fourth coordinate for time, we get analytical kinematics. One gets the
theorems of kinematics as theorems about invariants of a group of transformations of the
four-dimensional space of numbers. This is the group of Galileo.transformations. 20

Any physicist or philosopher of physics who wonders why these normatively


PROTOPHYSICS OF TIME 165

defined fundamental concepts of physics do not allow the successfully used


definitions of space and time of the special theory of relativity, at least as a
possibility, willieam

that the objections resulting from special relativity theory are already irrelevant here for
a trivial reason, i.e., because with the chronometry constructed here we are still always
concerned with a part of (haptic) protophysics independent from optical regularities in
the widest sense.
In contrast to that, an assertion can be advocated here which argues from the
established chronometry to special relativity theory: If the (... ) hypothesis of the
constant velocity of light in a vacuum and the associated definition of simultaneity at
separate places leads to the prediction that clocks moving relative to one another mani-
fest differences in their frequencies in a time comparison by means of light signals only
because of their motion, then from this the conclusion must be drawn that the compari-
son of clocks by means of light signals is an unsatisfactory procedure. In fact, no presen-
tation of special relativity theory is known which non·drculady establishes why the
attempt to produce undisturbed clocks, or at least to define them conceptually, was
given up (PoT, III, 4.33).21

And the protophysics of time finally culminates in polemic remarks against


the theories of relativity, for instance:

That Einstein's vindicated criticism of the completely unsatisfactory clarification of the


conceptual foundations of classical physics has led to the relativity theories must - on
behalf of the claim demonstrated here, that every scientific proposition needs methodi-
cal foundation - be called a historic disaster. It is not decided at all, whether phenomena
which cannot be described classically could not be explained by means of a small modi-
fication of the empirical part of classical physics. Historically the fascination which
without doubt arises from Einstein's considerations ... prevented the appropriate
improvement of the classical theory (PdZ, p. 107).

Because of the pretentious claims of proptophysics which lead to these


surprisingly polemic remarks against the theory of relativity, it would be
interesting to demonstrate at which particular place in the protophysical
approach towards a normative foundation of the basic physical concepts a
physicist is inevitably led to the fundamental concepts of Galileo-Newtonian
kinematics. Such a demonstration can best be performed if one follows the
development of the protophysics of time step by step and consciously
dispenses with any reference to concepts which do not belong to the class of
geometrical and chronometrical termini admitted by the protophysicists
themselves at this stage in the development of the protophysical program.
Examples of such forbidden technical terms would be 'propagation of light' ,
'transmission of signals', 'principle of the constancy of the velocity of light'.
166 JOACHIM PF ARR

Allowed are references to those mathematical theories which can be presup-


posed as completely or at least partially normatively founded, such as arith-
metic, Euclidean geometry and parts of analytical geometry. This way one
can achieve an immanent analysis of the protophysical program, for the
explicit reference to the tools of analytic geometry enables a reformulation of
Janich's verbally given basic definitions and propositions in a more formal
mathematical language without changing the meaning of the particular
termini. In order to achieve a better understanding, selected literal quotations
of some of the definitions in Janich's book are needed. The formal transcrip-
tion into the language of analytical geometry will be denoted as 'supplement'.
Since not all of Janich's definitions are needed for a critical analysis, a restric-
tion to the most relevant ones seems to be advisable and useful.

1. Straight Uniform Motion


Janich's principal goal is to answer the question: 'Suppose that Euclidean
geometry is well founded in the protophysical sense. What has to be added to
geometry in order to get kinematics?' Since 'change' and 'movement' can be
perceived in the pre-scientific world, the basic concept of clock-free kine-
matics is 'motion' in a pre-scientific sense: one thing moves (or is moved)
with respect to another. Besides the ability of man to move bodies with
respect to each other, he is able to guide bodies along given geometrical lines,
in particular, along straight lines (at least along parts of those). Thus by
analogy with the geometrical terms

line distance point

Janich introduces the terms:

motion process position,

and the geometrical traces of those 'motions':

trajectory path place.

'Motion' is now a technical term. A 'motion' has no beginning and no end.


'Processes' along definite paths are bounded by 'positions'; processes are
parts of 'motions' (PoT, III, 2.35).
After these definitions, Janich starts to compare motions, for simultaneous
PROTOPHYSICS OF TIME 167

motions (in the pre-scientific sense of simultaneity) can be compared with


each other. For the comparison of motions, lanich introduces several defini-
tions (see Figure 1):

/
/
/
/
/
/
K1

Fig. 1. The guide-line.

(Dl) Two point bodies KI and K2 move relative to Ko on trajectories B IO and B 2 0.


The intersection of the parallel to B 10 through K2 and of the parallel to B 20 through
K 1 describes a curve relative to Ko: let this curve be called 'guide·line 1 12 '. With
situations in which the motion relative to the same reference body has not been
stipulated for all point-bodies, let us write 1 12 • 0 to avoid misunderstandings. The index
located after the semi-colon indicates the refe~ence body. As a further restriction for
the following discussion, let the requirement always be fulfilled that the motions con-
sidered are 'smooth', i.e., they contain no reversal positions. Every place of a trajectory
will be run through only once (PoT, III, 3.35).

Supplement

The introduction of the guide-line in this manner obviously contains several


restrictions:
- The trajectories of the point-bodies Kl and K2 have to intersect at
the position of the reference body K o , which means that besides parallel and
skew trajectories, all those intersecting trajectories are forbidden which do
not intersect at the very position of Ko.
- The possibility of constructing a guide-line by means of intersecting
straight lines parallel to the trajectories of the point-bodies already presup-
poses a kind of naive concept of simultaneity. It must be - with respect to
168 JOACHIM PF ARR

Ko - 'simultaneously' known where the point-body K2 is, when the point-


body Kl is at the place S 1.
Since, due to the constructing procedure, the positions of the three bodies
K o, K 1 , K2 are all different, any reference to a 'simultaneity at the same
place' is not allowed. The guide-line 112;0 is deftned with respect to the
reference frame of the point-body Ko. Hence in this reference frame, the
possibility has to be given to acquire knowledge about the positions of the
two moving bodies Kl and K 2 , and this knowledge has to be given even if
the distances of the bodies cannot be measured by means of measuring
devices which originate from the normative foundation of Euclidean geome-
try (for instance, if the distances are large in comparison to those found in
pre-scientific experience). The use of signals of any kind fails here owing to a
logical argument: the form of motion of the signal propagation has to be
known in order to defme forms of motion.
Janich knows about these difficulties, yet instead of problematizing the
concept of simultaneity he refers only to "comparisons of motions which
can be carried out on 'simultaneous' (in the pre-scientiftc sense) motions"
(PoT, III, 3.40). Despite these fundamental difficulties which give rise to
severe deftnitional problems already at this stage of the deftnition of the basic
concepts of a clock-free kinematics, the discussion about simultaneity here
will be postponed to a later section.
In terms of analytical geometry, the deftnition for the guide-line can be
rewritten in the following way: any two intersecting straight lines spread
a plane. Each point of this plane can then be uniquely characterized as
an intersection of two parallels with respect to the initial straight lines.
Since Euclidean geometry can be presupposed with a metric (PoT, III, 3.58),
any point in the plane can be determined by means of a pair of real num-
bers (S1, S2). The gUide-line can be interpreted as a graph of a continuous
mapping

112 : R - R
(1)
S1 ~ S2 = 112(sd
which has to be monotonous due to the required 'smoothness'. And so long
as no one of the two moving point-bodies is at rest with respect to the other
one, strict monotony can be required, which includes that the mapping 112 is
reversible
PROTOPHYSICS OF TIME 169

From mathematical propositions about continuous functions, it follows that


a parameter representation can be given for each continuous curve in R2.
Here this means

Al12 :R "JM-+NCRV S: R -+M X NR2


(2)
8 1---+ (S 1 (8), S2 (8)) = S(8).

For the domains of definition and values the following relations hold for

M N
motions =R =R
processes ~ ~
positions {SI} {S2}

(The domain of the parameters [8 1 , 8 2 ] can be restricted to compact interval


without loss of information.) A parameter representation of this kind,
however, is not unique; there are infinitely many different representations
for the same guide-line. (See Figure 2.)

Ko ~
Fig. 2. Similar motions.

(D4) Kl and K2 move relative to Ko on non-parallel trajectories. Their motions b 10 and


b20 are called 'similar' if the guide-line 112 is straight (in symbols: b 10 * b 12 ) (PoT, III,
3.39).

Supplement

The use of straight lines as particular curves in R2 considerably restricts the


universality of the propositions about motions. Independent of the choice of
the parameter representation, the relation
170 JOACHIM PFARR

(3)

has to be true (kl and ko are constants with kl =1= 0).


Janich proves several propositions for 'similar motions', which can now
be reduced to propositions about parameterized curves in R2. However,
Janich's propositions are not necessarily needed explicitly, therefore it is
more useful to dispense with them and proceed to the steps towards an intro-
duction of uniform motion.
If one wants to define a particular form of motion by means of instruc-
tions for action, one has to be able to repeat processes in order to compare
the guide-lines with former results. Therefore the next important step is to
define what has to be understood by repeatable processes.
(D9) If repetitions of the motion comparison are begun respectively with the same pair
of concomitant positions and if there arises thereby the same guideline, then the com-
pared events may be called 'repeatable relative to one another'. If the operations of two
instruments G 1 and G2 , i.e., the events of their indicators Kl and K 2 , are repeatable
relative to one another, they may be called 'invariable relative to one another' (PoT, Ill,
4.17).

Supplement
In terms of the parameter representations two processes are called 'repeatable
relative to one another', when along with the parameter representation

S: [0 1 ,0 2 ] -MXNCR2
(4)
o 1-+ (SI (0), S2(0))
the following parameter representation holds

S : [0 1 + ljl, O2 + lj] -+ M X N C R2
(5)
81-+ (SI (8), S2 (8)) == (SI (8), S2 (0))
with lj > I O2 - 0 1 I and the same values (Sb S2)' In particular, this means
that the geometrical figures produced by the guide-line-construction for pro-
cesses which are repeatable with respect to each other are congruent.
The next restriction refers to (D4):
(DI0) Let two instruments G 1 and G 2 be invariable relative to one another. If the
running ratio of G 1 and G2 is constant, i.e., if the events of their indicators K 1 and K2
PROTOPHYSICS OF TIME 171

are similar, then Gl and G2 'agree with each other', and the events of their indicators
are 'repeatably similar' (PoT, III, 4.22).

Supplement
In addition to the parameterization with (D9) now the constraint

has to be imposed. (k 1 and k2 as in (D4).) This condition, however, still does


not suffice to produce a uniform motion. All motions with a constant ratio of
the velocities of the point-bodies Kl and K2 satisfy this condition. The next
definition leads to the wanted result (see Figure 3):

Fig. 3. Similar processes which are displaceable with respect to each other.

(D1l) Let two instruments G1 and G2 agree with each other. If, then, with repetition of
the operations without maintaining identical starting positions, a likewise equal and
constant running ratio exists, i.e., if arbitrary partial events of the indicators K 1 and K2
are similar to each other and have an equal momentary velocity ratio, then Gl and G 2
may be called 'equal to each other' and the events of their indicators 'displaceable
relative to one another' (PoT, III, 4.23).

Supplement
Now in addition to the parameter representations in (09) and (DlO), the
following holds:
172 JOACHIM PFARR

s: [01>0 2 ] -+MXNCR2 (6)


01-+ [Sl(0),S2(0)=K 1 S 1 (0 +r)+Kol
with arbitrary 'value of displacement' r.
A comparison between the two equations for the displaced motions

(7)

and

(3')

shows that they represent parts of straight lines in a (S 1 - S 2 )-coordinate


system. Hence

S2 (0) - S2 (0) = k 1 (S 1 (0 + r) - S1 (0)) = constant, (8)

and since r is arbitrary, the linearity of S 1 with respect to 0 immediately


follows 22 :

(9)

as well as the linearity of S2 with respect to 0:

(9')

If two instruments satisfy the conditions listed in the definitions (D9)-(D 11)
(which means that their indicators satisfy the linearity requirement with
respect to the parameter 0), then Janich calls these processes "regular" and
the motion which they describe is the "straight uniform motion" (PoT, III,
4.25).
And Janich can now say what a clock is: "(DI2) An instrument on which
a point-body moves uniformly is called a clock" (PoT, III, 4.26).
The parameter 0 can be interpreted as a time. According to Equation (9),
this time has the metric of the one-dimensional Euclidean continuum. And
the guide-line denotes the totality of all 'simultaneous' positions in the
coordinate system defined by the trajectories of the two moving bodies with
respect to this time. However, as already mentioned above, the time para-
meter () is not unique, although due to the restrictions to the motions given
PROTO PHYSICS OF TIME 173

by the definitions (D9)-(D11), there are some constraints as to the choice of


the parameter e itself. Thus before talking about time, the class of meaningful
parameter representations has to be found.
Introducing a coordinate system with equidistant intervals along the trajec-
tories of the motions of the point particles K 1 and K2 with respect to K o , the
motions can be written in the form

X==klle+kol (10)

(10')

where X and Yare the coordinates along the axes. (10) and (10') represent
equations of straight lines if e is interpreted as an additional coordinate.
Together with (10), (10'), all coordinate systems, (X, Y, 'if) are equivalent as
to the description of the straight uniform motion, which preserve the basic
properties of straight lines. Since the similarity of the motions of K 1 and K2
guarantees the linearity of, for instance, the motion along the Y-axis if the
motion along the X-axis is linear with respect to e, the restriction to a co-
ordinate system (X, e) is allowed.
The coordinate transformations which transform straight lines in (X, e)-
coordinates into straight lines in (X, 'if)-coordinates are the so-called collinea-
tions:
- aX+be+c iiX + be + c
X == AX + Be + C' 0 (11)
AX + Be + C·
(a, b, c, ii, b, c, A, B, C are constants.) Under these transformations, Equation
(10)

is transformed into

(11')

The requirement that the new coordinate system have the same reference
body as the old one, gives the conditions b = B = 0, the additional require-
ment that finite paths be transformed into finite paths gives A = 0 and C =1= O.
And if one dispenses with the pure coordinate translations, the fmal result is
a three-parameter family of transformations
174 JOACHIM PFARR

X=Ft·X
8'=FJ,o+Fl'x
with arbitrary constant parameters Fk (i, k =0,1). (12) transforms straight
uniform motions along the X-axis into straight uniform motions along the
X-axis. Besides the dependence on 0 to be expected, the new 'time'-parameter
8' depends on the position of the point-body in the old coordinate-system.
This means that the concept of 'simultaneity' is relativtzed: 'simultaneity'
of two positions X(O), YeO) with respect to the parameter 0 does in general
not imply 'simultaneity' with respect to the parameter 8, although the geome-
trical trajectories of the motions and the guide-lines are congruent.
From these results first of all the conclusion must be drawn that before
the straight uniform motion is declared to be the only correct reference
motion for the definition of clocks, a decision has to be made as to which one
of the possible coordinate systems has to be chosen as a basis for the concept
of simultaneity. This can be done by convention, since any reference with
respect to signal propagation is forbidderi. It could be done by means of an
additional instruction for action, which by itself would be of conventional
character, of course.

2. Straight Uniformly Moving Reference Frames

With regard to a kinematical theory, it is not sufficient to have one clock at


rest with respect to a particular reference frame. One has to know how
clocks behave when they are moved. Within the framework of clock-free
kinematics, this problem has to be solved by means of a reference to geome-
trical congruence propositions alone. (see Figure 4.) Again Janich uses pre-

Fig. 4. Def"mition of cpo


PROTOPHYSICS OF TIME 175

scientific experience: two point-bodies Kl and ~ are moved relative to the


point-body Ko on non-parallel trajectories B 10, BlO respectively. The
motions are assumed to be similar; the guide-line intersects Ko. Then it is
claimed by Janich that the triangles formed by the three bodies are congruent
in "simultaneous descriptions," no matter which of three bodies is chosen as
reference body, i.e., the body from which the other two bodies are "seen."
Janich recognizes the difficulty which lies in the concept "simultaneous
description," but instead of discussing this point he only claims: "To be sure,
the word 'simultaneous', used here for clearness, may be avoided as well and,
to that extent, does not violate methodological order" (foT, III, 4.43).
Using this pre-scientific concept of simultaneity, Janich, by analogy with the
rigid bodies in geometry, introduces what he calls "undisturbed clocks":
The rigid measuring body serves to compare lengths independent of the individual place
and the individual time of the comparison. Its peculiar property thus shows during
transport. Analogously the undisturbed clock has the task to give 'time comparisons' in-
dependent of the place and time of the comparisons. Their property consists of a still to
be described unchangeability (Unveriinderlichkeit) during transport through space and
time (PdZ, p. 105).

The unchangeability to be described means that any motion can by no means


affect the frequency of a clock. And, in the long run, this does mean the
introduction of an absolute time-coordinate with unique time-measurements.
As has been shown in the preceding sections, 'simultaneity' cannot be
uniquely defined in that frame of reference which is related to the point-
body Ko. 'Simultaneity' with respect to reference frames which move relative
to each other has no well-defined meaning at this stage in the development of
clock-free kinematics. And in the same sense, the concept of congruence of
geometrical objects in moving reference frames has no meaning at all.
Before such concepts as 'simultaneous' or 'congruent' are used in connec-
tion with moving reference frames, the problem of the transformation- equa-
tions from one frame of reference to another has to be solved. The solution
of this problem will be given in the next paragraph after a summary has be
made of the presuppositions which have hitherto been used in the protophysi-
cal program. Without reflection, Janich makes use of the following properties
of space, time and physical processes:
(1) The repeatability of processes in a pre-scientific sense already requires
a time topology with present, past and future. It must be understood what it
means when two processes which are developing now, develop in the same
way later. likewise the knowledge must be acquired about processes which
already have developed earlier.
176 JOACHIM PFARR

(2) Comparisons of past, present and future processes by means of con-


gruence properties of geometrical figures alone can only be made meaning-
fully if the durations of the different processes are indistinguishable by them-
selves. That is to say, time has to be homogeneous.
(3) Processes as finite parts of motions develop in three-dimensional space.
The concepts of 'repeatability' and 'replaceability' do not distinguish by
themselves special places or directions in space, i.e., space is assumed to be
homogeneous and isotropical (with respect to the defmition of motion and
forms of motion).
(4) Singular processes, 'seen' from reference frames moving straight
uniformly relative to each other, are indistinguishable as far as their forms of
motion are concerned.
For the protophysicists, all these presuppositions belong to pre-scientific
experience and are therefore not reflected upon by them. These propositions
can, on the other hand, be interpreted as the conditions for the possibility of
performing the instructions for action, which are needed to define, for
instance, clocks. (If, for example, the homogeneity of time is not warranted,
the protophysical procedure fails at the problem of 'replaceability'.) In
content, these propositions are equivalent to a particular version of a set of
propositions known in physics by the name 'principle of relativity'. This
principle reads (in an active formulation): from a frame of reference 10 the
course of physical events is indistinguishable (a) at any place in space, (b) in
any direction of space, (c) at any time, (d) in any frame of reference I which
moves straight, uniformly with respect to 10 •
lf we introduce Cartesian coordinate systems (x, y, z, t), (x', y', Z', t') in
10 and 1 respectively, then - as was first shown by Frank and Rothe in 1911 23
- the coordinate transformation between these two reference frames under
the conditions (a) to (d) is
1
x' = (x - v· t)· (1 - v2 /v;,)- 2
1
t' = (t -v ·x/v;') . (1 - IJ2 /IJ'!f 2 (13)
y' = y, Z' = z.

This transformation is called the Generalized Lorentz- Transformation. v is


the velocity of 1 with respect to 10 ; IJ~ is an undefined parameter with
the dimension of a velocity. In order to make this transformation unique,
one has to impose an additional requirement. This could be, for instance,
that the propagation of light is isotropic in each of these coordinate-frames
PROTOPHYSICS OF TIME 177

but it could be, of course, the demand that clocks are 'undisturbed' and that
triangles seen' from different frames of reference are congruent. The former
request would lead to the well-known Lorentz transformations, the latter one
to the Galileo transformatiOn. Yet this latter demand would not automati-
cally include that mechanics is Galileo-invariant. It would only stipulate one
special coordinate system and not an invariance-property of nature. For even
if one particular value for v~ is fIxed by means of a conventional requirement
or by means of some invariance found in the behavior of nature, there is still
the possibility of choosing particular coordinate systems in each of the
reference frames. To show this explicitly, one has to investigate the
coordinate transformations allowed within one frame of reference which do
not violate the principle of relativity.

3. Gauge Transformations in Fixed Reference Frames

Up to the present, the two additional space directions have been neglected
when coordinate transformations within a given frame of reference were con-
sidered. This was justifIed, since once the motion of a point-body K 1 with
respect to the reference body Ko was known to be straight uniformly, the
corresponding motion of the second body had this property by means of the
similarity of the motions. Now since the problem of congruence arises, at
least one additional space coordinate has to be included. The makes an
extension of the admitted coordinate systems with respect to the spatial
coordinates necessary. 24
In particular, the two reference frames defmed by the two point-bodies
Ko and Klare chosen for the analysis. (In the following, the reference frames
will, for simplicity, be denoted by the denotation of the point-bodies them-
selves.) In both Ko and Kl there is a seven-parameter transformation of the
type

(x, y, z, 8) T(F) (x, y, i, lJ) (x,' y,, Z,, 8')----+(-' -, Z,


T(j) x, y, -, 8-')

x = F~ .x + F~ . y x' = fi . x' +f~ . y'


y = F1·x+F~·y ji' = f1 . x' +f~ . y'
i =z i' = z'
lJ = Fg·8+FY·x+pg·y 0' = fg . 8' + fY . x' + fg .y' (14)

as a generalization of the three-parameter family (12). The requirement that


178 JOACHIM PFARR

the new coordinate systems can be connected by means of a generalized


Lorentz transformation of the type T(v, v~) yields the conditions

Ii = Pi = 0, ~ = f4.
The additional requirement that the coordinates along the axes of motion
are independent of the coordinates of the other axes (an assumption already
following from the isotropy of space) reduces the coordinate transformations
to the four-parameter families
-, ,
X F~ ·x x = .11 . x
-, ,
ji=~.y y = f4 . y
-, ,
i z z =z
{J = fo . 0' + fl . x' (15)

with the relations

Fl = FUv-~ v fl = fo/v - IUv.


Furthermore, the relations
1 1
.11= vlv(1 - v2Iv~ ft . . (1 - v2/I)~r 1.. ~
1 1
fo= vlv(l- 1)2/1):')2· (l-v2Iv;'r2·F~

hold between the parameters in the two different reference frames. Hence

T(j) in K 1 depends only on one independent parameter I:,


it follows that, given a particular set of parameters in Ko, the transformation
for instance, fo.
The special choice Fi =F g =~ = I, Fl = 0 in Ko leadS to the set of para-
meters in K 1 (v =I) for this special choice). 2 5

fo = (1 - v2Iv:')!. (l - v2Iv:.r! = (11)-1


fl = - (1 - (f8)2)/(vfo) (16)
The interpretation of this one-parameter family of coordinate transforma-
tions within a given reference frame is the following: even if the value of v~
has been fixed by means of some invariance property of nature of by means
of an additional conventional requirement, it is still possible to introduce a
PROTO PHYSICS OF TIME 179

coordinate system in KI which represents a different generalized Lorentz-


frame and thus does not violate the principle of relativity.
These results can now be applied to the congruence problem of the
triangles 'seen' from different reference frames.
Let the motions of the three point-bodies, K I and K2 with respect to K o ,
be described by
Xo =0 Xl = V' 0 X2 =0
(17)
Yo = 0 Yl = 0 Y2 = u'O
v and u are the velocities of the point-bodies K I and K2 with respect to K o ,
respectively. In the reference frame of KI these motions read in the co-
ordinates x', y', 0'

Xo' = -J). 0' X2' = -v' 0'


, 2 2 I ,(18)
Yo' =0 Y1' = 0 Y2 = U' (l -v Iv.. )'. . 0 ,

and for the angle <p' in the primed coordinate frame, is obtained

tgcp' = tg<p (l - v2 Iv:')~ . (19)

Using the coordinate transformation (15) with (16) in K I, the equations of


motion read

xot = -J). 8 X2' = - l ) . 8'


I _
(20)
Yo' = 0 :P2' = u·(-v2Iii:')"z 0'

and the angle in the new coordinate system is given by


I
tg<p' = tg<p' (l _v2 Iii;') 2. (21)

It was assumed that the considered triangles are rectangular ones. Thus
congruence is given if one side and an additional angle are equal. The
common side Ko Kl is always warranted, however, the equality of the angles
is given only if in Kia coordinate system with ii.. = 00 is chosen. This co-
ordinate system is connected with the original coordinate system in Ko by
means of a Galileo transformation
-, -, {}' = O.
x' = x-v' 0 Y =Y z = z (22)
180 JOACHIM PFARR

One can obtain this result by just putting v~ = 00 in the generalized Lorentz
transfonnation (Equation 13), of course. However, it was important to show
that independent of the specific value of v~ there is still freedom in the choice
of the coordinate system in each reference frame which could lead to Galileo
transfonnations. Yet this freedom in the choice of coordinates must not be
mixed up with the existence of genuine invariance properties of nature.
TIlls proves the statements given at the end of the last section, that
Janich's required congruence of triangles 'seen' from different reference
frames, only stipulates the use of Galileo coordinates and does not automati·
cally imply the corresponding invariance properties. In the same sense
Janich's 'undisturbed clocks' have to be understood: the requirement of
undisturbed clocks only means the use of a Galilean time e as derived above
and thus is merely connected with the corresponding choice of the coordinate
system.

4. The 'Undisturbed Clocks' in Relativistic Kinematics

Due to these coordinate representations, it is possible to define 'undisturbed


clocks' conceptually and give instructions for their construction even in the
framework of special relativity, i.e., even if the parameter v~ is fixed to v~ = c
(and c is the velocity of light in a vacuum). TIlls can easily be seen by means
of the following example.
Let 10 and l' be two inertial systems,Io moving with respect tol' with the
velocity v = -315.c. The two reference frames are connected by means of the
Lorentz-transfonnation

X'= (x+ v' t)(1_V2/C2t~


t' =(t+ v ·x/c 2)(1- ,,2/C2t~

In both reference frames there are equally constructed clocks at the positions
Xi and x/ (i € N) respectively. These clocks are assumed to be synchronized in
each reference frame by means of light signals (Einstein synchronization) or
slow-clock-transport. 26 The problem now is to synchronize the clocks in l'
with those in the particular frame Io.
The coordinate transformation

x' .t1. x' with.t1 = 4/5, .fo 5/4, If = 3/4


i' = to . t' +If . x'
PROTO PHYSICS OF TIME 181

transforms from Lorentz-coordinates to Galileo-coordinates with the frame


of reference t. From the coefficients f ~ it can be seen what kind of construc-
r
tive features are necessary to synchronize the clocks in with those in 10:
(1) The clocks have to go faster for 25%, since ",i' Ix' = '" t' .1& = 5/4 . "'t'.
This property can be achieved by readjusting the balance of a normal clock.
(2) Since the time coordinate t' depends on the positions x' of the clocks
in t, all clocks synchronized with respect to t' have to be put on for the time
-If ·x'. This can be done by a readjustment of the indicators.

Direction of Motion •

Fig. 5. The 'undisturbed clock'.

A model for such a clock is shown in Figure 5. A screw serves to adjust the
balance to the correct frequency of the clock (here 25% faster). Furthermore,
the clock stands on wheels and one of these wheels is connected to the indi-
cators by means of a transmission. The farther the clock is moved away from
the origin of I' in positive x' -direction, the faster it is, and vice versa. This
clock is a model for Janich's 'undisturbed clocks'. In x'-direction this clock
has to be transported with rolling friction; motions with respect to the other
space directions must not influence the speed of the clock, which means that
182 JOACHIM PFARR

the orientation of this clock with respect to the spatial axis has to be pre-
served. Moreover, this clock has to be transported so slowly that a further
time dilation of relativistic origin due to the motion of the clock with respect
to t does not occur. Yet this clock is always synchronized with those in the
special reference frame 10 .
It is pOSSible, of course, to use measuring devices of that kind for physical
experiments. However, the above considerations show one of the reaSons why
physicists will dispense with the use of instruments based on the normative
prescriptions of protophysics: there are no dynamically operating
mechanisms which automatically force clocks to be 'undisturbed' if kinema-
tics as a physical theory turns out to be not invariant under the group of
Galileo-transformations.
There are other reasons why physicists would reject the proposals given
in the protophysical program. Some of them will be discussed in the next
.section.

III. PROTOPHYSICS VERSUS PHYSICS

As is well known and is often complained about by the protophysicists,


physical theories are not established according to the protophysical proposals.
Therefore it would be interesting to compare the two different approaches on
the basis of a simplified exemplary confrontation on the question 'how does
one pass from Galilean to relativistic kinematics? Neglecting historical,
linguistic and - in some respects - methodological aspects in the develop-
ment of the special theory of relativity, it will only be pointed out how it
would be possible, in prinCiple, to detect the Lorentz-invariance of nature of
means only of mechanical devices.
Let us start with the physiCist. Physicists will - at least in the respect that
we need to consider here - agree with protophysicists that physics as an
empirical science is first of all based on pre-scientific experience and
knowledge. In this 'life world', constant behavior of nature in the repetition
of processes and invariances (in a pre-scientific sense) can be realized and kept
in mind (for instance, the changes of day and night, the seasons, flooding of
rivers, falling of apples from trees, constant timing of a pendulum). Among
these processes there are some which seem to reflect the invariance properties
and constant behaviors perfectly well. These processes are then used for an
operational definition of the basic concepts (here, for instance, the astrono-
mical time as 'time' and the motion of the stars as 'clock'). The 'artificial'
clocks then have to coincide with this clock. Using an appropriate definition
PROTOPHYSICS OF TIME 183

of length (by means of solid bodies), one can formulate Galilean kinematics
as a physical theory and this theory can be experimentally tested by means
of measuring devices which are derived from the operational definitions. For
the present, it is expected that those invariances are reproduced which have
been constitutively used in defining the concepts and in constructing the
measuring devices. TIris is always the case, as long as the experimental
conditions do not exceed the restrictions taken into account when the
concepts were defined and the measuring devices constructed. In our concrete
example, this would mean that, as long as the velocities we want to measure
do not exceed those found in the life world, the experimental results will con-
firm our theory and the assumption of an absolute time. Suppose now that
the experimental conditions are made such that the velocities leave the
domain of velocities out of which the conceptual constituents for measuring
devices have been taken, and for which the theory has been formulated.
There are two possibilities:
(1) The theoretical predictions tally with the measurement. This would
mean that the theory is valid even for an enlarged domain of experience, and
with the theory, the basic concepts and measuring devices, of course.
(2) The experimental results deviate from the theoretical predictions. The
'undisturbed clocks' turn out to be disturbed. First of all, this would indicate
that the theory is not made for this part of experimental experience. But it
would say in addition that the basic concepts and measuring devices need a
review, since the invariance properties of nature exhausted for the definition
of concepts and devices, tum out to be not the real invariance properties of
nature. So one has to look for the new invariances, define new basic concepts,
defme new measuring devices whose construction features can be the same as
formerly. The clocks of Galilean mechanics can be used in relativistic
mechanics too. Only their interpretation as measuring devices differ. One
then proceeds to the experimental confirmation of the new theory and its
new basic concepts. Thus, if there is an invariance in nature that differs from
that one found in the life world, the physicist with his operationally defined
measuring devices could find this invariance. And in addition, he would try
to describe nature using the properties of the new invariance.
What does the protophysicst do? He starts from the same pre-scientific
conditions, but he will construct his measuring devices normatively on the
basis of his life world experience. He will produce 'undisturbed clocks' which,
with logical necessity, will remain undisturbed as long as his experimental
conditions remain in the domain of the life world experience. If the experi-
mental conditions leave this domain and there is another invariance property
184 JOACHIM PFARR

of nature which cannot be realized in the pre-scientific world, he, with his
instruments, of course, would measure these effects in the same way as the
_physicist did (although this statement is still doubted by Janich27). But
instead of revising-his conceptual system and the conceptual meaning of his
measuring devices, he would try to describe the new effects in the old
concepts. In particular, he would describe Lorentz-invariant properties in
Galilean coordinates. All this is possible and does not lead to logical con-
tradictions. Whether it is useful or not is a different question.

theory theory

operational constructions
definition af - scientific of measuring scientific
measuring - experience devices experience
devices Protophysics
f
pre-scientific experience I
pre-scientific experience

Fig. 6.

The main differences between the two approaches obviously lie in the way
of defining basic concepts and measuring devices and in the different inter-
pretation of the role of pre-scientific experience. In physics, the possibility
that operationally defined basic concepts might be inconsistent with the
results of scientific experience is provided in the methodological establish-
ment itself and can be regarded as a desirable element with respect to scien-
tific, normatively defined basic concepts and measuring devices. As a cons-
sequence, all scientific experience here has to be formulated in 'pre-scientific
language'. As already pointed out, in some cases this is possible and logically
free of contradiction. However, one must not expect that measuring devices
'eo ipso' behave according to the norms founded in pre-scientific experience
when the experimental conditions leave this domain. Measuring devices as
objects of nature succumb to the invariance principles which are inherent in
nature whether they are normatively defined or not. From the first, clocks
will not automatically behave undisturbedly in the protophysical sense.
However, by means of some constructive changes and additional apparatus, it
is possible to force clocks to behave like that. But in order to achieve this,
firstly one must pay special attention to one particular frame of reference
(where this additional equipment is not necessary) and secondly, there are no
PROTOPHYSICS OF TIME 185

dynamically operating mechanisms which could automatically force clocks


to be 'undisturbed'. These considerations clearly show the high price one has
to pay for sticking to pre-scientific insights of invariance principles relative to
which an empirical physics turns out not t6 be invariant. By consciously
renouncing the regulative feedback between scientific experience and the
operational definitions of basic concepts and measuring procedures,
protophysics complicates the approach towards domains of scientific
experience and theory establishment which could be more easily approached
by means of a less restrictive operationalism.
We return to the suspicion expressed before that the ostensibly pre-
suppositionless propositions about reality in fact do depend on the conditions
for the possibility of performing the instructions for action. We could ex-
plicitly formulate some of these conditions and show that under these con-
ditions the protophysical approach to kinematics and chronometry is only
one possibility, but that it is not a necessary one. There are other possibilities
as well, especially that of relativistic kinematics, and therefore the polemics
against the theories of relativity are not justified at all. In particular, this
relativistic possibility - although no longer founded in the life world - can
be made an empirical scientific experience even if only such measuring
devices are used as are exclusively constructed according to the protophysical
prescriptions. This demonstrates that pre-scientific experience can only be an
incomplete basis for an empirical science.
We can thus state that it is the interpretation of the stock of experience in
the life world as a priori which exposes parts of the protophysical program to
the danger of running counter to empirical results. Interpreting the '0 priori'
in this less restrictive sense, the protophysicists have - expressed again in
Kant's terminology - made propositions about the conditions for the possi-
bility of the objects of experience without making sure that these conditions
are likewise the conditions for the possibility of experience itself.

Institut fur Theoretische Physik, Universitiit zu Koln

ACKNOWLEDGEMENTS

The author is pleased to thank Professor L. Laudan for a critical reading of an


earlier version of the manuscript. He is very grateful for the hospitality
extended to him by Professor A. Griinbaum and the University of Pittsburgh.
186 JOACHIM PFARR

NOTES

* This work was performed when the author was a research associate to the Center for
Philosophy of Science at the University of Pittsburgh, Pittsburgh, Pennsylvania.
1 P. Janich (forthcoming), referred to within the text as PoT followed by the relevant
chapter and section numbers.
2 A collection of partially reprinted papers can be found in G Bohme, 1976. This book
contains essays written by P. Janich, F. Kambartel, P. Lorenzen, and J. Mittelstrass as
advocates for a normative foundation of physics, as well as papers by G. Bohme, W.
Buchel, K. J. Diisberg, A. KamIah and P. Mittelstaedt as opponents and critical com-
mentators on this program. Reprinted papers by different authors concerning the rela-
tions between protophysics and the theory of relativity are also collected in J. Pfarr
~ed.), 1981.
Janich, 1969, referred to within the text as PdZ followed by the page number. (The
translations are my own. Under the same title, P. Janich, 1980, published the new
German edition of this book. The new volume is revised and contains many replies and
discussions concerning critical comments made by various authors.
4 I. Kant, 1787, B. pp. 19, 20ff., 128 and 874. See also I. Kant, 1783, p. 295.
5 I. Kant, 1787, B, pp. 224 and 232.
6 Ibid., B. pp. 221 ff.
7 Ibid., B. p. 197.
8 H. Dingler, 1923, p. v.
9 Ibid., p. 36.
10 H. Dingler, 1921, p. 114.
11 H. Dingler, 1964, p. 42.
12 H. Dingler, 1974, pp. 77ff.
13 For a more detailed analysis and survey of the Dinglerian system, see P. Janich,
forthcoming (PoT, II, Hf.).
14 P. Lorenzen, 1968b, p. 128.
15 P. Lorenzen, 1967, p. 60.
16 P. Lorenzen, 1976, pp. 5-6.
17 P. Janich et al., 1974, p. 89.
18 Ibid., p. 89.
19 A suspicion of this kind was first formulated by P. Mittelstaedt in his (1975), pp.
38ff. Mittelstaedt's claims, however, and his interpretation of the conditions for the
possibility of performing the instructions for action differ from those presented here.
Mittlestaedt guessed that these conditions are logically equivalent to the allegation of the
validity of at least some of those propositions which can be proved from the normative
deilnitions. This conjecture is justified and correct if the conditions for the possibility
of performing operational constructions are interpreted as, by themselves, being strictly
related to pre-scientific experience. Here the conditions for the possibility of performing
instructions for action are presented in the less restrictive sense, and it will be shown that
under this less restrictive interpretation, these conditions allow for more than opera-
tional constructions based on pre-scientific experience alone.
20 P. Lorenzen, 1976, pp. 5-6.
21 'Undisturbed' clocks are clocks which by no means change their frequency during
transport through space and time (cf. P. Janich, 1969 (PdZ p. 105». A more de-
PROTO PHYSICS OF TIME 187

tailed discussion of undisturbed clocks is given in Sections 11.2 and 11.4 below.
22 The linearity of Sj (j = 1, 2) with respect to (J is easily seen if Equation (8) is read in
the form
Sj (u«(J» - Sj«(J) = constant, with u ((J) = (J + 7
and differentiated with respect to (J:
dS. dS.
du ' .
du
<IlJ - dlr = 0, (A)

or, because of the special form of u«(J)


dS j = dS j .
du d(J
The left-hand side is a pure function of u, the right-hand side the same function of (J:

dS j = : fi«(J) = fi(u) and u = (J + 7,


d(J
this condition can only be satisfied if Ij is a constant. Hence
Sj«(J) = Ii . (J + k oj •
It is important to mention that for this derivation, the homogeneity of the 'time'-para-
meter (J has been needed (Le., intervals [(Jb (J2) can be shifted along the (J-axis and
are congruent to the corresponding intervals upon this axis). If the (J-axis is allowed to
be inhomogeneous, (J in (8) has to be replaced by some 9, 9
where is a nonlinear func-
tion of (J. The condition (A) then only reads
dS. du dS. 0
du'·d(J - d(J' = ,
with u = 9 + 7, which is a mere consequence of the chain-rule for differentiable functions
and always holds independent of the particular way in which Sj depends on (J. Trans-
ferred to the problem of the definition of 'uniform motion', this would mean that a
uniform motion without the assumption of a homogeneous time could not be defined
at all by means of geometrical arguments alone.
23 P. Frank and H. Rothe, 1911. See also E. Wiechert, 1911, H. Arzelies, 1966, Section
53; Y. P. Terletskii, 1968, Section 7; and G. Sussmann, 1969. For a more recent repre-
sention, see 1.-M. Levy-Leblond, 1976, and P. Mittelstaedt, 1977, p. 154.
24 A more detailed analysis of the properties of these coordinate-systems, which are
possible within a given frame of reference, has been given by C. Mflller, 1972, p. 267, or
can be found in H. Heintzmann and P. Mittelstaedt, 1968, Section 1.
25 The denotation is taken from P. Mittelstaedt, 1976.
26 See, for instance, A. Grtinbaum, 1973, Ch. 20, and the literature quoted there, or
H. Reichenbach, 1958. Ch. 19.
27 W. Buchel in his (1976), p. 241, argues that "two protophysicists moved relative to
each other would find out by themselves about the Lorentz-invariance of the laws of
nature". (Translation my own.) lanich, in his reply (Janich, 1976, p. 310), rejects this
thesis by means of a reference to the "uniqueness of the homogeneity principle" for the
uniform motion.
188 JOACHIM PFARR

REFERENCES

Arzelies, H. 1966. Relativistic Kinematics. Oxford: Pergamon Press.


Balzer, W. and KamIah, A. (eds.). 1979. Aspekte der physikalischen Begriffsbildung.
Braunschweig: Vieweg.
Boiune, G. (ed.). 1976. Protophysik: FUr und wider eine konstruktive Wissenschafts·
theorie der Physik. Frankfurt: Suhrkamp.
Biichel, W. 1976. 'Zur Protophysik von Raum und Zeit.' In Protophysik, ed. by G.
Bohme. Frankfurt: Suhrkamp.
Dingler, H. 1921. Physik und Hypothese. Berlin: de Gruyter.
Dingler, H. 1923. Die Grundlagen der Physik. Berlin: de Gruyter.
Dingler, H. 1964. Aujbau der exakten Fundamentalwissenschaft. Munich: Eidos.
Dingler, H. 1974. Die Ergreiftmgdes Wirklichen, ed. by F. Kambartel and J. Mittelstrass.
Frankfurt: Suhrkamp.
Frank, P. and Rothe, H. 1911. American Journal of Physics 34 825.
Griinbaum, A. 1973. Philosophical Problems of Space and Time, 2nd, enlarged ed.
Boston Studies in the Philosophy of Science, Vol. 12. Dordrech t, Holland: D. Reidel.
Heintzmann, H. and Mittelstaedt, P. 1968. 'Physikalische Gesetze in Beschleunigten
Bezugssystemen.' Springer Tracts in Modern Physics, Vol. 47. Berlin: Springer Verlag.
Janich, P. 1969. Die Protophysik der Zeit. Mannheim: Bibliographisches Institut.
Janich, P. 1976. 'Zur Kritik an der Protophysik.' In Protophysik, ed. by G. Bohme.
Frankfurt: Suhrkamp.
Janich, P. 1977. 'Die Protophysik der Zeit und das Relativitiitsprinzip, Erwiderung auf
Joachim Pfarr.' Zeitschr. f allgemeine Wissenschaftstheorie 9 343.
Janich, P. 1980. Die Protophysik der Zeit: Konstruktive Begriindung und Geschichte der
Zeitmessung. Frankfurt: Suhrkamp.
Janich, P. Forthcoming Protophysics of Time. Boston Studies in the Philosophy of
• Science, Vol. 30. Dordrecht, Holland: D. Reidel.
Janich, P., Kambartel, F., and Mittelstrass, J. 1974. Wissenschaftstheorie als Wissen·
schaftskritik. Frankfurt: Aspekte.
KamIah, A. 1979. 'Zur Diskussion urn die Protophysik.' In Konstruktionen versus
Positionen (Festschrift for Paul Lorenzen), ed. by K. Lorenz. Vol. 1, pp. 311-339.
Berlin: de Gruyter.
Kant, I. 1783. Prolegomena to any Future Metaphysics, translated by L. W. Beck.
Indianapolis: Bobbs-Merrill, 1950. (Translation of Prolegomena zu einer jeden
kiinftigen Metaphysik. Riga, 1783.)
Kant, I. 1787. Critique of Pure Reason, translated by N. K. Smith. London: Macmillan,
1929. (2nd edition of Die Kritik der reinen Vernunft, Riga, 1787.)
Levy-Leblond, J.-M. 1976. 'One More Derivation of the Lorentz-Transformation,'
American Journal ofPhysics 44271.
Lorenzen, P. 1967. Normative Logic and Ethics. Mannheim: Bibliographisches Institut.
Lorenzen, P. 1968a. Methodisches Denken. Frankfurt: Suhrkamp.
Lorenzen, P. 1968b. 'Das Begriindungsproblem der Geometrie als Wissenschaft der
rliumlichen Ordnung' In Methodisches Denken (see above). Frankfurt: Suhrkamp.
Lorenzen, P. 1976. 'Discussion of the Four Fundamental Units of Measurement and
Their Defmition.' Translation of 'Zur Definition der vier fundamentalen Messgrossen.'
Phil. Nat. 16 1.
PROTOPHYSICS OF TIME 189

Moeller, C. 1972. The Theory of Relativity, 2nd edition. Oxford: Clarendon Press.
Mittelstaedt, P. 1975. 'Der Wissenschaftsbegriff in der Physik.' Studia Leibnitiana,
Sonderheft 5.
Mittelstaedt, P. 1977. Der Zeitbegriff in der Physik, Chapter VII, p. 127. Mannheim:
Bibliographisches Institut
Mittelstaedt, P. 1979. 'Protophysik der Zeit und Spezielle Relativitatstheorie.' In Kon-
struktionen versus Position en ... Vol. 1, pp. 290-310.
Pfarr, J. 1976. 'Die Protophysik der Zeit und das Relativitatsprinzip.' Zeitschr. f allgem.
Wissenschaftstheorie 7 298.
Pfarr, J. 1979. 'Zur Eindeutigkeit der Zeit in der Protophysik'. In Aspekte der physikali-
schen Begriffsbildung, ed. by W. Balzer, A. Kamiah, p. 147.
Pfarr, J. (ed.) 1981. Protophysik und Relativitiitstheorie. Mannheim: Bibliographisches
Institut
Reichenbach, H. 1957. The Philosophy of Space and Time. New York: Dover.
Siissmann, G. 1969. Z. f Naturf 24a 495.
Terletskii, Y. P. 1968. Paradoxes in the Theory of Relativity. New York: Plenum Press.
Wiechert, E. 1911. Phys. Z. 12689 and 737.
PETER JANICH

COMMENTARY ON 'PROTOPHYSICS OF TIME AND THE


PRINCIPLE OF RELA TIVITY'*

O. OBJECTIONS TO PROTOPHYSICS

In the lively discussion of protophysics of the last ten years, two types of
objections have been raised: (a) immanent objections that accept the program
of protophysics, but point to mistakes in the execution as presented, and
(b) external objections. These can in turn be roughly divided into (lb)
philosophical objections (until now only made from the position of modem
empiricism) and (b2) physical objections (by means of a confrontation of
protophysics and physics). In J. Pfarr, all three types of objections can be
found, although his main interest lies in a critique of protophysics from the
vantage point of relativistic physics.

1. PFARR'S THESES

(a) Pfarr's immanent critique asserts that a naive concept of simultaneity is


used in Protophysics of Time in order to define clocks (and thereby simul-
taneity). This objection of there being a circular argument is, however, due
to a logical error on Pfarr's part.
(bI) Pfarr's philosophical objection asserts that Protophysics of Time does
not reflect the empirical conditions for the possibility of normative proto-
physics. Therefore the protophysica1 norms are not meaningful or appropriate.
This objection, however, is due to a methodical circle, that results from the
fact that Pfarr does not distinguish between the linguistic presuppositions of
propositions and the nonlinguistic conditions of actions.
(b2) Finally, Pfarr's physical objection contains a twofold mistake: in
order to make the assertion that the invariances of nature are covered by
the Lorentz transformations, as well as for his objections to protophysics
connected with it, he himself first has to acknowledge a protophysics. Thus
he no longer has the possibility of refUting it. Furthermore, the 'Pfarr clock'
is no model of the protophysical definition of clocks.
In order to justify these assertions the core of the protophysical program
will be introduced, since Pfarr does not even mention, much less take into
account, the crucial idea of this program, i.e., the principle of methodical
191
R. S. Cohen and M. W. Wartofsky (eds.), Physical Sciencerand History ofPhyliics, 191-197.
© 1984 by D. Reidel Publishing Company.
192 PETER JANICH

order. 1his leads to the consequence that Pfarr confronts physics with a
caricature of protophysics rather than with protophysics itself.

2. PROTOPHYSICS AND KANT

Whether the methodical philosophy of physics as developed by Dingler,


Lorenzen and lanich follows Kant's theory of the synthetic a priori is a diffi-
cult question, especially of Kant-interpretation. It will not be considered
here since protophysics does not rely on Kant as an authority. Rather
protophysics can be understood without any reference to Kant. Pfarr also
does not derive any arguments from Kant, he merely inverts the famous
formulation borrowed from Kant, from 'a priori conditions of the possibility
of empirical propositions' to 'empirical conditions of the possibility of a
priori propositions'. Why this reversal with regard to protophysics is
meaningless, can also be shown without reference to the authority of Kant.
Therefore the question whether protophysics is a priori will not be important
in what follows.

3. THE BASIC IDEA OF PROTOPHYSICS

Physics as a modern science of experimentation cannot get by without


measuring devices. Measuring devices (e.g., clocks) are artifacts. As such they
have artificial qualities, i.e., those systematically produced by man, as well
as qualities incidental to production. If we limit ourselves to the considera-
tion of those qualities that make an object a measuring device and disregard
other 'incidental' qualities (e.g., in the case of clocks the speed of the hands
as opposed to the color and weight of the clock), then the measuring device
is distinguished by intended qualities, which the producer of the device seeks
to induce and the user of the device seeks to maintain. The purposes of a
device (i.e., what a device has to achieve in order to be a measuring device,
e.g., a clock) are at the same time the aims of production and are formulated
in protophysics as norms for production or use. Such norms are to be under-
stood in such a way that they guide the actions of producers or users of
measuring devices in the crucial points.
Obviously it is not the same to establish or acknowledge a norm and to
satisfy a norm successfully. Thus there is a difference between establishing a
norm for the uniformity of the movements of the hands (and thus to define
'clock') and constructing a clock successfully. Even ifit is known how clocks
are supposed to work, they sometimes do not work well or not at all. There-
COMMENTARY ON 'PROTOPHYSICS .. .' 193

fore there is with regard to the relevant qualities, Le., those defining the
measuring device, yet another distinction of artificial qualities (namely, in
the case of clocks, those achieved by the skill of the watchmaker) and natural
qualities (Le., those occurring spontaneously). The former can be determined
as set purposes, thus making competent production or use of measuring
devices possible, the latter can be determined empirically as deviations from
the former. If, for example, a clock stops, because its battery is dead, then
this is a relevant, but natural, quality in the sense that the watchmaker or
physicist does not pursue this as a purpose nor does he attempt to achieve it
through action. Thus the natural qualities always occur as deviations from the
intended behavior of measuring devices and thus, by definition as well as in
their empirical determination, they remain dependent on the qualities as
measuring devices that were determined as stated purposes.
It is, however, false to view measuring devices exclusively as natural
objects by attempting to comprehend their function as solely empirical-
physical - as does 'Pfarr. ("Measuring devices as objects of nature succumb to
the invariance principles which are inherent in nature whether they are
normatively defined or not" (p. 184). Since defects are also events that can
be determined empirically/physically, a physical description alone does not
produce a distinction between disturbed and undisturbed clocks. Rather,
defects ('disturbances') are only understandable as deviations from the modes
of functioning of measuring devices seen as normatively determined and
technically produced.

4. THE RELATIONSHIP OF PROTOPHYSICS AND PHYSICS

If a physicist cannot distinguish between undisturbed and disturbed func-


tioning of his measuring devices in the laboratory, he does not obtain
measurement data with empirical content. (An extreme example: all events
at a place are 'simultaneous', when the 'clock' used has stopped. Physicists
are, however, only interested in temporal measurement data that are obtained
by clocks that' operate undisturbed.) Protophysics inquires into operative
definitions for the undisturbed operation of measuring devices.
The point that distinguishes protophysics from analytical theories of
measurement is the principle of methodical order. This principle consists
of the prohibition of logical, definitional and pragmatic circles. While there
is a general attempt to avoid logical and definitional circles, this is not always
the case with pragmatic circles in theoretical physics. Above all, Pfarr ignores
this prohibition. Statements about the sequence of actions, which can not
194 PETER JANICH

be carried out in the stated sequence are called pragmatically circular, because
the results of 'later' actions have to be available as the condition for allegedly
'earlier' actions. Pragmatic circles are thus only possible in speech, in theory,
but not in action, or in practice.
With this in mind, Pfarr's objections can now be discussed and rejected.

S. PFARR'S IMMANENT CRITICISM

Pfarr accuses me of having "presupposed" "a kind of naive concept of simul-


taneity" (p. 167) in two places. Where this accusation concerns the definition
of a guide-line, he does not recognize that a guide-line is created by a machine
with parallel straight lines (rods) for which geometrical determinations only
are sufficient. Here the methodical sequence of steps is important: the term
'simultaneity' is not needed in order to formulate directions for the produc-
tion of a machine in which a guide-line is produced for the purpose of the
comparison of motion without clocks. Once this machine is defmed (and
produced), however, then 'simultaneity' of (mechanically coupled) events can
be defined by means of it as a second step.
At another place (p. 175), Pfarr repeats his accusation and objects to the
formulation "simultaneous description". Pfarr overlooks that 'simultaneous'
does not appear here in a temporal, but in a logical sense: neither physics nor
protophysics nor a clock is needed when it is stated that, together with a state-
ment A, a statement B that is logically equivalent to A is asserted 'simul-
taneously' - and that we are dealing inProtophysics of Time not with a tem-
poral, by with a logical equivalent which is exactly what is demonstrated.

6. PFARR'S EMPIRICAL PHILOSOPHICAL CRITICISM

In his philosophical objections, Pfarr makes the mistake of not distinguishing


between the linguistic presuppositions of statements and the nonlinguistic
conditions of actions, here, of protophysically established (normierten)
actions. He talks about "hidden presuppositions" and "unreflected presup-
positions" or "the conditions for the possibility of performing the proto-
physical instructions for action" (p. 164) and then cites as examples state-
ments from physical theories.
He who does not equate the reality, in which physicists practice physics,
with the reality about which physical theories make statements, and who
then mistakenly confuses these with the theories themselves, will understand
that the possibility of actions can only be checked by acting. On this level
COMMENTARY ON 'PROTOPHYSICS .. .' 195

Pfarr does not make objections because the fulftlment of protophysical


modes of action would hardly pose difficulties for him. Therefore his objec-
tions are to be understood in such a way that he already has theoretical
knowledge from physics at his disposal that argues against compliance with
protophysical directions for action. It would be absurd to imagine that one
would have to know first of the homogeneity of time or the homogeneity and
isotropy of space, i.e., explicitly to have formulated and established these
as statements, in order to be able to execute the actions of the construction
of measuring devices. (Watchmakers of even the best modem clocks manage
without these theoretical statements.) What would it therefore mean then, to
say that homogeneity and isotropy are supposed to be conditions of acting,
of, say, the production of clocks? Actions themselves, however, are not state-
ments (and only statements can follow logically from other statements and
thus have 'necessary' presuppositions). Thus homogeneity or isotropy of
space or time are not presuppositions of actions established (normierten) by
protophysics. Therefore it is false when Pfarr insists, "Janich makes use of
... properties of space, time and physical' processes ..." (p. 175). No objec-
tion can be made, then, that the analytically true statement about proto-
physical norms is made subsequently, namely that protophysics leads to a
homogeneous time, but this analytical knowledge presupposes practical
compliance with the protophysical norms through actions.
With the accusation that protophysics is unreflective with regard to its
own "presuppositions", Pfarr documents that he has misunderstood proto-
physics as a whole, because he ignores the principle of methodical order. the
whole and probably the only point of protophysics lies exactly in reflection
on conditions of actions and presuppositions of propositions. A non-circular
foundation of physics opens up the possibility of argumentation for or against
physical statements aided by presuppositions in that the 'first' propositions
that are 'presupposed' as definitions are tied to the nonlinguistic, teclmical
production of devices. These production processes for their part are prescribed
by instructions and do not have linguistic presuppositions with a truth value.

7. PFARR'S PHYSICAL CRITICISM

Pfarr's main physical objection amounts to the accusation that protophysics


wants to force qualities of invariance on measuring devices which are not
those of nature. Protophysics clin~ to old norms (i.e., norms leading to
classical physics) - and thus to classical physics itself - contrary to better
empiricit1-physical knowledge.
196 PETER JANICH

If the "invariance properties of nature," cited by Pfarr several times, are


not a metaphysical monster, it must be that invariances of empirical proposi-
tions of physics in transformations are referred to. These propositions as
empirically valid propositions are based on measurements, at least as far as
their claim goes. At this point Pfarr overlooks the difficulty of his argumenta-
tion: he either does not recognize protophysics, and then the claim to validity
of the cited physical propositions on the basis of measurements is not
redeemable, or Pfarr recognizes protophysics, and then it is irrefutable by
empirical theories because of its normative character.
(It should be noted here that Pfarr does not develop a theory from his own
suggestions (p. 184f.), which could compete with protophysics. He does not
say what operational definitions that can be 'revised' are supposed to look
like. To choose "the motion of the stars as 'clock' " (p. 182) may be useful
for everyday use, but it is not an operational definition of clocks. Without
measuring devices 'artificial' clocks cannot even be used as simulators of
natural processes - quite apart from the fact that arguments for the choice
of the motions of the stars aided by physical theories are again circular.)
Pfarr's physical criticism of protophysics suffers from the fact that he
reformulates parts of protophysics in the framework of physical theories and
as physical theories. The result is meaningful neither for physicists nor for
protophysicists. This is shown above all by his construction of clocks, which
surely is not a model "for lanich's 'undisturbed clocks' " (p. 181). For it is de-
fined within the framework of relativistic kinematics. Relativistic kinematics,
however, as an empirical theory, presupposes a measurement of time with
undisturbed clocks. No physicist would use Pfarr's clock, and no proto-
phYSicist would recommend its use. Pfarr's clock is probably a scientific joke.

8. SUMMARY

Pfarr's· critique of protophysics on the whole amounts to the argument that


the impossibility of protophysics follows from physical theories. If that be
so, protophysics must be interpreted as a partial theory of physics, which it
is not. Pfarr disregards the methodical, the operational and the normative
character of protophysics. Therefore he misunderstands not only the proto-
physical program and its execution as a whole, but also the protophysical
critique of relativistic physics, the differentiation of the epistemological
status of Galileo and Lorentz invariance, and the methodical difference
between a protophysical and a physical proposition. He overlooks the
difference between speaking and acting as well as their presuppositions and
COMMENTARY ON 'PROTOPHYSICS ... ' 197

conditions. Expressed pointedly, Pfarr does not speak at all about the proto-
physics of the protophysicists.
Pfarr's critique can also be viewed as follows: he makes physics the
epistemological theory of protophysics. For the protophysicist this leads to
meaningless statements because they are circular. What, however, if circu-
larity were not prohibited? Then there still remains the problem of the
epistemological foundations of physics. What would be the sense for the
treatment of this problem in confronting physical theories with theories of
different content and with a different claim? Pfarr says nothing about that.

Philipps-Universitiit Marburg

A NOTE ON THE LITERATURE

Several times, Pfarr cites: P. Janich, Die Protophysik der Zeit (Mannheim: Bibliogra-
phisches Institu t, 1969). Since then there has been a profusion of objections which were
published together with my reply in G. Bohme (ed.), Protophysik: Fur und wider eine
konstruktive Wissenschaftstheorie det Physik (Frankfurt: Suhrkamp, 1976). Further-
more, also published was: P. Janich, Die Protophysik der Zeit: Konstruktive Begriindung
und Geschichte der Zeitmessung (Frankfurt: Suhrkamp, 1980). The English translation
is based on this latest text. In it there is a chapter on the critique of protophysics -
including Pfarr's critique. Therefore I do not understand why Pfarr, who knows the
English text claims that my book "will be published in English without a critical com-
mentary". Indeed, both the German and the English versions contain a bibliography of
all critical commentaries to date.

* The author is grateful to Renate Hanauer for translating this paper.


PAUL M. QUAY, S. J.

TEMPORALITY AND THE STRUCTURE OF PHYSICS AS


HUMAN ENDEAVOR *

Wir leben nicht unmittelbar im Sein, daher


wird Wahrheit nicht unser fertiger Besitz.
Wir leben im Zeitdasein: Wahrheit ist unser
Weg. (Karl Jaspers, Von der Wahrheit, p. 1.)

Those philosophers who, over the last two decades, have been studying scien-
tific discovery, the history of physics and its themata, research traditions,
'scientific revolutions' and, in general, the temporal aspects of physics, should
not be seen, I think, as turning aside from the more rigorous and difficult
domain of philosophy itself. Rather, the degree to which physics changes in
time and, perhaps especially, its penchant for holding onto and continuing to
use outmoded theories, though knowing them to be in contradiction to newer
ones, has become an increasingly serious problem for philosophy. Such
changes and contradictions seem incompatible with any claims which physics
might make to truth, even to the lower-level truths of consistency and of
accuracy of the data-base called for by logical positivism. Given the temporal
flux which is physics, can one still speak meaningfully about knowledge,
through physics, of the empirical world? And if physics gives no truth, can
any science do so?
To sharpen the problem slightly, we may note that what has been sought,
largely under the influence of physics, or even presumed to be possessed, in
the period since Galileo - so largely dominated by the Cartesian problematic
- has been 'the perfect method', some sort of automatic defense against
error. This was not always taken to mean that the possessor of the method
would be rendered proof against mistakes, but, at least, that the method
itself would guarantee the correction of all mistakes if one would continue
to utilize it. Logical positivism, no less than the older approaches to philoso-
phy of physics, sought within physics itself, as a method, the grounds for its
success. Such a one-sided approach, of course, had ultimately to eliminate
the notion of truth. The newer approaches have the merit of having burst out
of this constrictive Cartesian framework. Seeking to understand physics from

199
R. S. Cohen and M W. Wartofsky (eds.), Physical Sciences and History of Physics, 199-230.
© 1984 by D. Reidel Publishing Company.
200 PAUL M. QUAY

the point of view of the individuals, groups, or communities doing it, they
have become aware that whatever physics may have attained in the way of
truth is best grasped from a study of its history, since even 'revolutionary'
physics grows out of a tradition and retains more of its past than it discards.
It is useful to recall Galileo's image of nature as a great book written in the
language of mathematics, requiring of us only that we learn that language in
order to read it. The object of physics, so presented, is not primarily the
description of the vagaries of physical objects in time but the 'reading' of
what is there permanently written, i.e., the discovery of the transtemporal
laws that govern all changes. As a result of the successes of Galileo and those
who followed him, physics came to be seen as the ultimate objective
knowledge, that is, as the basic and only truly adequate description of a know-
able world other than our minds, as the only non-tautological knowledge
independent of the concrete psych!Jlogical states of individual knowing
subjects. Its truth, rooted in the objects known, was thought to be capable of
being made evident to any mind whatsoever which had the same background,
the same data, the same attitudes and attentiveness.
At the same time, physicists and philosophers, both, began to see our
knowledge of the world as some sort of fusion of empirical data and mathe-
matics, sliding from that into an attempt to construct objective knowledge
from a fusion of what most deserves the designation 'subjective' (our sen-
sations produced on the occasion of meter-reading;) with what is mere tau-
tology (according to their interpretation of mathematical propositions). No
wonder that Holton reports despair among intellectuals, a surmise that there
is nothing real at the center of this maze (Holton, 1973, pp. 30-36).
The paradox has been strengthened in recent decades; physics as a body of
objective knowledge - perhaps the most perfect of all in its balance of em-
pirical and theoretical, its drawing in orderly and effective ways upon mathe-
matics and logic, history and philosophy - seems marvellously uncorrelated
with physics as the art of discovering such knowledge. (By 'art' I mean a dis-
cipline which uses feeling, 'intuition', the non-rational, for the achieving of a
rational goal which does not, however, wholly determine the means to be used
in its achievement.) PhysiCS as objective knowledge, which Holton has called
"public physics" (Holton, 1973, pp. 17-22 and 387-391), is accessible to
any mind whatever and the same for all, tested for just this quality by the
community of physicists. The art of physics, Holton's "private physics", is
subjective, compact of passions, idiosyncrasies, personal feuds, joys and frus-
trations, errors, even theology and 'mysticism'. But how can the randomness
of pure subjectivity be the source of the universal necessity of objective truth?
TEMPORALITY 201

Kuhn, Laudan, and many others, have pushed the matter further by
emphasizing the changes which have taken place in the objective content of
physics, noting that these changes - and not only during "scientific revolu-
tions" - have been inextricably tied to subjectively motivated decisions.
Historical studies have made increasingly evident that the nature of the disci-
pline itself, its purposes, and its methods, no less than its conclusions, are all
profoundly time-dependent and vary greatly with the individual physicist, his
culture, and his own psychological conditions. "Public physics" seems, like
foam on the ocean, to float in patches whose shape and structure are only
slightly more permanent than the swirls and breaking waves of feeling and
desire that stirred them up. As a result, these philosophers have simply set
aside the possibility that physics can arrive at truth, that it can know any-
thing of this world other than the mind of the physicist. 1
In this paper, though, rather than respond directly to any of the problems
mentioned, I should like to sketch a synthetic framework, arising from the
way in which physicists in fact do their work, which may enable us to deal
not only with those problems but with many others.
Most, perhaps all, of the difficulties objected to physics-as-knowledge can,
I think, be adequately met by means of an epistemology less naive than those
usually brought to bear on the matter. Nevertheless, the temporal approaches
that have been seen as dissolving the truth of physics suggest a different way
to come at the matter, a way with its own special advantages. I shall argue
here that in all its aspects - as knowledge, as method, as discipline - physics
is intrinsically, not just psychologically, temporal. Each aspect of it is, in its
own way, a type of activity of intrinsically temporal beings, interacting in a
temporal flow of history in a world changing in accord with laws which may
themselves, for aught we know, be changing. Yet, I shall urge, it is possible
to find an epistemological significance to these temporal aspects of physics,
and the most profoundly temporal element of all is that which guarantees
for us the possibility (not the fact) of truth: the free exercise of our power to
choose.
It has long been known - this was indeed one of the primary motive-
forces behind the Cartesian and subsequent epistemologies, to go no further
back - that error comes from the human affectivity: from free choice at
times no less than from passion, emotion, feeling, the unconscious. What has
not been enough appreciated is that the formal recognition of truth is im-
possible without the intervention of freedom, that intellectual knowledge is
impossible without something very much akin to love. The physicist's
temporal activities themselves are of epistemological worth and involve truth
202 PAUL M. QUAY

at all stages and in all their varieties. The ability to appreciate such truth
depends essentially upon his freedom. Hence, knowledge of truth is essen-
tially incompatible with the possibility of an infallible method.

II

Such divisions of physics as those into public and private, subjective art and
objective knowledge, temporal and transtemporal, whatever usefulness they
may have in certain contexts, are seriously lacking if taken as adequately
dividing the field. Whatever is pointed to as objective knowledge, as subjective
drives or experiences, as public information or private ponderings is all
included, along with much else, in 'physics' in the sense I shall be using here,
that is, the entire range of those particular human activities through which
we know or seek to know, in a methodically secure way, how material
objects behave, along with explanation of such behavior ip. terms of their
ultimate constituents and of the laws governing their motions and interactions.
Now, such activities are, always and only, the activities of individual
physicists. Knowledge, no matter how objective or public, is a modification, a
property or a quality, indeed an activity, of some knowing subject who is an
individual with his own individual senses and mind. If the knowledge is
public, this means that others also know it or have it accessible to them and,
usually, that each knows that the others have it. But physics is not a substan-
tive entity and does not exist apart from some human mind. If there are no
human minds (and, likewise, none of the imaginary minds which give
Popper's arguments on this point their plausibility (Popper, 1973, pp. 115-
116) - no minds of putative extra-terrestrial visitors, no mere ghostly projec-
tions of our own into distant futures), there is no physics. Should too many
neutron bombs be exploded, the world, insofar as a subject-matter for
physics, presumably remains unchanged. The books and journals which con-
tained its data and laws might also remain. But if no one is left to read these
journals and to reconstitute their signification in some mind, physics no
longer exists. It is, then, always and in all its aspect'>, a mode of human
activity: a knOwing, theorizing, writing, observing, arguing, evaluating, meas-
uring, debating, remembering, proving, constructing, and the rest, 'Physics'
is our abstract terminology for this struggle and endeavor, along with all that
knowledge which we gain through it. Hence, anything we say of physics can
be said, more concretely, of the activities of physicists.
If all this is painfully obvious, yet it is too easily forgotten. Nor is it
unimportant for the problem of relating physics as art and as knowledge. But
TEMPORALITY 203

even were physics to be restricted solely to knowledge, we would, if not tying


that tightly to the activity-of-knowing of single individuals, still be dealing
with what does not exist. For, physics is far too vast in scope and too
complex at all its levels to be grasped by any single mind. PhYSics, then, as a
field, is a mere being of reason, existing nowhere. No one knows it. What we
refer to in so speaking is the entire range of what 'in principle' any mind
could know at the present moment in this doinain, an abstract aggregation of
the concrete knowledge of each. There is more to say on this matter; but I
should like to defer it for now and begin to look at physics from the point of
view ofits temporality.

III

Analysis of temporal relationships can be a refmed tool for a phenomenologi-


cal study of the types of human activity which characterize the physicist as
such. Though time enters this activity in a quasi-infinity of ways, these can be
grouped quite naturally into relatively few classes, which in tum can show us
something of the inner structure of the physicist's endeavor. Our own
temporal situation in this talk, tonight, however, hardly permits any pheno-
menological developments. Let me simply state some of the most obvious
results of such an analysis, with specific regard to scientific theories, and then
comment on a few points of special interest.
The most obvious classes or types of temporal change are probably the
following: (1) the changes occurring empirically in the objects which a
physical theory is set up to describe;2(2) the characteristic ways in which
a theory itself undergoes modification; (3) the development and growth of
the theorizer, whether as constructor of the theory or as student ofit; (4) the
processes taking place both within the community of those who are able to
use or validate theories and in relation to the society in which it grows; (5)
the variations in the notion of what a theory is and what is expected from it.
A. Now, a few comments. Firstly, time affects physics in the sense that
physicists have made time an integral portion of their subject-matter. Most
of physics is concerned with the description, prediction, measurement, and
calculation of time-dependent phenomena. Hence, time appears as a variable
in the laws and theories of physics, as a measure of the changes occurring in
the world representatively described by physics. All these temporal features
of the world, however, are transformed by conceptual analysis and theore-
tical structuring into intrinsically non-temporal laws and theories. It is
through these highly prized, transtemporal, cognitive portions of physics that
204 PAUL M. QUAY

we seek the power to describe all possible physical processes taking place in
time.
B. Yet the laws themselves are continually being changed by physicists -
through minor extensions, restrictions, or generalizations, or else through
complete overhaul or replacement. This fact: that the cognitive configuration
of physics changes, brings us up squarely against the historicists' problem: 3
If physics is true knowledge, how can it change in time save by accretion or
restructuring? Without seeking an answer yet, let me divide and recast the
question.
Firstly, we are not diSCUSSing whether the laws of physics themselves might
not prove intrinsically time-dependent. If there is a temporal variation of our
laws of physiCS and if there are, also, more remote laws governing them, then
the ultimate freedom from temporality would be found in those remoter
laws. Were there no such superior laws, physics would be impossible as an
exact science over any sufficiently long period. Yet our laws as formulated
could still be valid, in whatever degree of approximation, over some suitable
neighborhood of the present moment.
Secondly, there is no doubt but there are strictly non-temporal aspects to
our thoughts. The conceptual forms we receive, possess, or construct, while
they may change psychologically as they come and go in our minds or are
altered by forgetfulness or deliberate reconstruction or else by a change of
our valuation of their usefulness in bringing our minds into conformity with
reality, have in themselves no intrinsic mutability. For example, the pure
forms of classical mechanics remain today what they were when first
formulated by Newton, Euler, and Cauchy. Their formulations, exemplifica-
tions, and applications have changed. They have been subordinated to other,
more· comprehensive theories. Most importantly, we have revised our
judgments about their adequacy for describing the world. For all that, those
forms are as conceivable today as when invented, even though some historical
investigation might be necessary to guarantee their identity.
We no longer think in terms of phlogiston, still less of an elemental fire.
Yet these notions, whenever had, are what they were. Our frequent loss of
ability even to think them is a psychological matter, usually based upon the
thought that there is nothing in the world to which they correspond, so that
any judgment asserting their existence or activity in our world is necessarily
false. Concepts and theories, then, when reflexively inspected as mental beings
in their own right, involve no temporality in their form and structure. Even
our ideas of flows and alternating currents, of mutability and change, and of
time itself have no greater intrinsic time-dependence than any others.
TEMPORALITY 205

A true judgment is another kind of immutable, intelligible structure. This


remark holds, of course, only of that judgment understood as when made,
i.e., in its full, albeit momentary, context. Nor can such a judgment
legitimately be lifted out of that context save in virtue of some further truth
which justifies the generalization. Evidently, in a particular language such as
English, the verbal formula used may need to be changed in order to preserve
the original sense. 'I am now in Boston' is true. There will never be a time
when it will be falsified as to its content-taken-in-context. But, a week hence,
I shall not be able to express it save in terms of a past tense and a date. In the
same manner, a false judgment is forever false - for the truth is that, as
uttered, it was false. Physical laws, mathematical structures, and the pure
forms oflogic are similar 'cognitive' or 'formal' invariants.
Perhaps what has most obstructed recognition of the precise modes of
temporal-transtemporal dynamic in the thought of physicists is the concep-
tion of physics as a sort of Kantian Idea, either an imperfect one, the whole
field of present knowledge, mentioned earlier, or the ultimate and perfect
representation of the structure and laws of the physical world, sometimes
pictured as known by an equally non-existent human-mind-as-such, or
sometimes as an independent "world of objective knowledge", to use
Popper's term (popper, 1973, pp. 106-118, esp. p. 111), a knowledge un-
associated with any knower whatever and equally disjoined from the real
world of which it is knowledge.
Now, 1 have no wish to deny all usefulness to so conceiving physics. While
incapable of existence, such a fiction of the mind can stir strong desire. As a
goal, it can draw us on fiercely. But, in the last analysis, only this unattain-
able goal itself would be wholly transtemporal as a perfect intellectual form
of knowledge.4 Our danger is to think that, insofar as physics reaches truth, it
is nothing but an ever growing collection of bits and pieces of that Idea,
themselves equally transtemporal, intellectual fragments of subsistent know-
ledge. But if physics is nothing but an entirely and perfectly objective, even if
fragmentary, representation of the world, then how can it have any non-
objective aspects? Obviously it cannot; the question, so posed, is trivial; but
the actual doing of physics becomes an enigma. But if we avoid that trap, we
are free to ask - and at least begin to answer - a much more reasonable
question: How are the purely rational and objective aspects of physics related
to the rest of the physicist's activity?
The often drastic ways in which the apparently objective content of
physics changes do not require, however, that we reduce the transtemporal
aspects of our science to a psychological flux nor yet that we save them from
206 PAUL M. QUAY

reduction, either psychologistic or materialistic, by splitting them off as a


world apart. Rather, time has its principal and most direct effects upon the
relations between the components of intellectual structures and between
these structures and the world. Only perfectly simple concepts, wholly
lacking in internal structure, could be wholly free of this kind of temporality.
Even the bond of concept with language is enough structure of permit some
kind of change.
So, when a new idea is introduced, older ones are often reconceptualized.
This may be done by limitation. Thus 'quantity of heat' was split by Clausius,
who introduced the new concept 'entropy', leaving 'heat' much restricted in
application and much more useful. Or one may create an equivalent expres-
sion for the old idea in terms of the newer one. So, the heat fed into a system
in some process which takes it from one equilibrium state to another became
the difference between the work done when taking that system adiabatically
between those states and the work done on it in the actual process. More
careful analysis discovers other modes of variability.
As I have described elsewhere (Quay, 1974a), the internal growth and
modification of laws and theories never stops. Newton's theory was, for
better than 200 years, eminently successful and without serious challenge.
Yet it was continually transformed, tightened, augmented, extended, restruc-
tured, applied, and refined by d' Alembert, Laplace, Cauchy, Hamilton,
Poincare, to say nothing of its continued reworking by Truesdell and his
collaborators in today's 'rational mechanics'.
Theoretical methods also change. Some physicists will be alert for new
models and new ways of picturing what is going on; others will be busily
moving in the opposite direction with new ideals ofaxiomatization or
mathematical representation. New branches of mathematics are introduced
into physical analysis, usually piece-wise, for dealing with those problems
which seem more likely to be tractable by the new method - or simply
because the physicist happens to know that area of mathematics well. Then,
after an extended period, the whole theory may be recast in the new forma-
lism.
Nor is such change a matter only of concepts and theories. Just as our
theories are required to be equally valid for any mind whatever, so must our
data be. The movements of one's instruments need to be intellectually pro-
cessed even to be noted down. The mere, shifting blacks and whites of
tracings, emulsions, or meter-dials do not themselves tell us much, certainly
nothing that the editors would let into Physical Review. Nor does the exact
experience I have in reading a dial, the emotions that accompany it, or the
TEMPORALITY 207

color of my meter have any place at all in my data - yet all these are among
what were, in fact, given. Raw sensa, were such possible for human beings,
could never serve as data. The so-called theory-Iadenness of our empirical data
is but belated recognition of the fact that data are not, as such, sensibles but
intelligibles. Yet our data are evidently in even greater flux than our concepts.
Now, not merely will data-gathering, the running of experiments, have a
history. The description of the experimental system, the things done to it and
with it, the reliability of the readings, with estimates of random and systema-
tic error along with indications of data-spread, the theoretical analysis of the
linkage of the data with the effect to be measured, in brief, all the elements
of a very complex network of theory, of prediction, of both kinds of estima-
tion, go into the establishing of a single datum-point. The data finally
recorded as 'obtained' have not merely been measured and processed in time
but remain subject to changes in the relations among the elements of that
network (as well as to changes within these elements) and, thus, perpetually
open to revision.
New data enter into an already existing structure. 'Old' data are gathered
in ever new ways; new modes of processing them are invented; they shift in
their mutual relations due to the new facts, or else change. in significance
because we have altered our judgment about the property being measured or
used in the measurement. New empirical regularities are constantly turning
up, some sought for, some not. New problems call for greater accuracy, but
also require that one correlate and process the data differently in accord with
new theoretical questions. Thus, the use of Onsager's "reciprocal relations" in
processing thermochemical data, legitimate enough ordinarily, would vitiate
the data for the purposes of someone seeking to re-think the validity of these
relations on the basis of a new theory or in search of one. More profoundly,
as Heraclitus noted long centuries ago, no experiment can be performed twice
under identical circumstances, not even twice in immediate succession. The
earth spins and goes around the sun, itself moving off towards Vega in a
galaxy not fixed with respect to other galaxies. And this says nothing of the
ever changing conditions of body and mind of the experimenter.
Much attention has been given recently to changes in the problems
addressed by physicists. Though a problem exists, as such, only for the person
who is in fact puzzled by it and though the psychological processes involved
are extremely fluid and complex, I think that problems, also, are cognitive
invariants with the same basic sorts of changeability as those already
discussed. As with those, so the elements of a problem shift within it as one
works on it and seeks to bring it into focus. Or else, it changes through its
208 PAUL M. QUAY

ever shifting relations with other parts of physics and, as Laudan rightly
argues, with the rest of our intellectual activity. 5
Thus a problem can grow in time, becoming deeper, more complex, less
tractable as earlier investigations only uncover further difficulties, whether
resolving the initial one or not. Consider only the history of Boltzmann's
seemingly intuitively simple and straightforward question about how to carry
out a physically sensible averaging over phase-space. Other problems change
from merely empirical or experimental anomalies intg problems of basic
principle as formerly, for example, the photoelectric effect or the vanishing
of specific heats at low temperatures.
On the other hand, a major problem can, without being in any degree
resolved, become either trivialized or made to vanish as non-existent by a
suitable change in perspective. Galileo's great contribution was, with Des-
cartes' help and, ultimately Einstein's, to abolish entirely as a problem the
question: What is the cause of continuing rectilinear motion? In early
quantum theory, the question of what electrons do in the transition between
energy-levels in an atom was regarded as a non-problem and declared out-of-
bounds. Yet, once the theory had developed in some fullness on that basis,
this same problem was resuscitated and neatly solved.
I have written at length (Quay, 1975) concerning the often overlooked,
directly deductive aspects of discovery in physics, as exemplified in the
dialectical dynamic involved in the making and verification of predictions
and the correlative operations of mensurative and reconstructive estimation,
through which new data are gathered and theory is interrelated with experi-
ment. 6 Here, I would add that, beyond that complex temporal structure, is
the further, yet fundamentally similar process of direct reflection upon 'the
state of the discipline' - the reflective intercomparison of what is already
present in the way of data, regularities, laws, and theories. These are set off
against one another and related to all one's background knowledge in a quasi-
systematic attempt to detect something of the structure of what is still
unknown. This structure, evidently, must be continuous with that part of
reality which is already known and which, therefore, serves as a boundary-
condition to be met by any putative new knowledge.
This short list of ways in which the cognitive invariants of physics change
is by no means exhaustive. At least it may suffice to make clear that our
knowledge does change in time, not only by accretion of new information or
by borrowing from other changing disciplines such as logic and mathematics
but through continually shifting ways of understanding. Nor have I even
touched upon changes resulting from the loss of much that has been learned,
TEMPORALITY 209.

especially the less highly formalized and more poorly articulated components
of earlier insights.
C. Closely tied to the various changes in physics-as-knowledge are those of
our fourth category, communal changes within the community of physicists
and between that community and the larger society of which it is a part.
Thus, an unresolved divergence of viewpoint concerning the appropriateness
of some alteration of method or theoretical structure often results in the
formation of different schools or research traditions. Of greater importance
is the developmental process which leads to the dominance of a new, 'scienti-
fic' epistemology in physics and often beyond, It was such a transition that
separated Galileo and his successors so decisively from medieval physics.
Subsequent shifts, from those initiated by Descartes to those born of the
quantum theorists and of Einstein can each be shown to grow naturally
enough from earlier, not always widely accepted, epistemological changes.
Not surprisingly, a major epistemological shift, when successful socially, leads
physicists to slough off, where they can, the physics that was characteristic of
the earlier epistemology. I need not mention how strong an impact such
epistemological shifts in physics have had on the rest of the contemporary
culture, even if in ways which are traceable more to publicists and philoso-
phers than to the physicists.
Further, since this world is, in large measure, objectively knowable and
physics aims at just such objective knowledge about it, one hallmark of our
activity is that its results be intersubjectively knowable and testable. Without
a community to scrutinize his methods and results, a physicist would be
deprived of one of his most basic helps in his efforts toward truth. Key ideas
and great theories are, indeed, the work of single minds: GaIileo, Newton,
Euler, Maxwell, Planck, Gibbs, Einstein - no matter how truly these may be
said to synthesize, to rearrange, to winnow what has preceded them. Yet were
there no community of physicists, none of these could have developed. Nor
would there have been those capable of testing, applying, correcting, vali-
dating the work of such giants. For, giants in physics cannot appear as such
save to the eyes of those who understand them and who have themselves
wrestled enough with at least similar problems to be able to recognize the
victories achieved. More fundamentally, physicists are as much social beings
as the rest of mankind; and time will affect physics through every alteration
of its social embodiment. .
Yet this social aspect of our work has often been pictured in terms of the
labors of coral animals, each free-swimming awhile, then drawn back to build
upon the work of others and in interaction with them. The image has some
210 PAUL M. QUAY

merit. There is something like an exoskeleton for physics. Though it is neither


knowledge nor any other activity of physicists, it grows directly therefrom,
through the writing of articles and books, the constructing of schools and
libraries, laboratories and research centers, and the creation of those applica-
tions which ground technolOgies. Such artifacts let a community of physi-
cists transmit their understanding of the world to others with whom they
have no direct contact, especially to those of future generations; serve to
make their ideas visible, if not always intelligible,
/
to the rest of society; and
gain public import through their interaction with other societal groupings.
Yet the coral-image suggests that the growth, whether of objective know-
ledge or of the community, consists solely of 'more of the same'. It quite
fails to capture such social phenomena as 'scientific revolutions' or the
epistemological development that is characteristic of physics' growth. What-
ever societal equilibrium there may be in physics for a time is a dynamic one,
the social averaging of many actions and attitudes. And even in the quieter
seasons of physics' growth, such equilibrium is usually slowly shifting.
D. More slowly and deeply yet, what physics is thought to be as a cogni-
tive endeavor also is contingent and historically changing. This fifth way in
which time affects physics is often closely related to the second class above,
yet is at root distinct. The changes appearing there could, in principle, take
place without any change in the notion of what physics is and does. But, in
fact, this notion - I refer to the discipline, not the word - has a considerable
history. What began with the investigation of the motions of inanimate
bodies, from a largely philosophical point of view, changed gradually in its
object as concepts such as velocity, force and inertia, field, and energy
developed. Perhaps the most recent changes of our ideas of the nature of the
field occurred as determinate structure and morphic invariance entered
through quantum mechanics, making it possible to take chemistry firmly into
this generalized physics as a sub-field.
What constituted legitimate methods also changed from mere reflection
upon general and ordinary observations to the Aristotelian analysis of those
simple experiments in which there is less of control than avoidance of com-
plication - one does not so much make things do what is desired in order to
discover more about them as let them behave as they will with a minimum of
other factors present - and to the late medieval and Galilean thought-experi-
ments. There followed, gradually, experimental prediction and testing, with
ever increasing mathematization, an ever more sophisticated constructing of
physical and other kinds of models and analogs, arriving finally, with Newton,
at the construction of plausible world-views, and thence to the sophisticated
TEMPORALITY 211

self-reflections of the type of Mach and, rather differently, of Einstein and


Bohr, whence to the bold attempts to make the forms of physical laws them-
selves be temporally dependent on the current state of a particular universe.
Such changes of method have probably not come to an end.
Very similar remarks could be made concerning· what are taken to be
legitimate problematics for physics and the appropriate range of themata (to
use Holton's term). Yet, changes in our notion of what physics is, what
methods it may use, what kinds of problematics or themata are acceptable, all
shift the intellectual context for any individual's work and the sorts of formal
'tools' which he will have available to him. Such contextual shifting, while
evidently non-determining, certainly modifies the average 'shape' of the field
as it develops. It is in describing these changes, I think, that SOciology of
knowledge could make a legitimate contribution if it chose.
For example, the a priQri possibility of physics requires in us a certain kind
and development of our own consciousness and self-awareness. If we had not
yet learned how to depersonalize the world (for, some recent philosophy not-
withstanding, the great problem is not how to know or recognize other
persons but its converse, how to know and recognize the non-personal beings
of the world), then whatever regularities are observed would have become the
effects only of the choices made by animating spirits of the world.
Even with that step taken, no known primary oral culture has been
capable of physics (save in the simplest sense of presuming the future to be
like the past) since it lacks the elaborate abstract categories which make
systematic thought and long-sustained argument possible. Without records,
without even isolable words or reflexively grasped grammatical structures,
without tables or graphs, how could data be garnered and kept undistorted
long enough to serve as groundwork for an empirical generalization - how
indeed could it even be thought of as data, a notion that only began to make
its appearance in strongly cheirographic cultures?
The introduction of writing opens the way to phYSiCS, but essentially as
only a natural philosophy. It is not mere accident that Copernicus came 75
years after Gutenberg, and Galileo a good century and a half. A consciousness
begotten of exact reduplication, of visual iterability in space, of significant
pagination - the localization of knowledge - when combined with other
modes of visualization of the 'contents of the mind' deriving from mercantile
growth and the quantification growing therefrom, made it possible to think
of our experience in ways not formerly conceivable. The shift over the past
couple of generations from the dominance of the media of print and static
image to those of exactly iterable visual flows - we seem as yet to have no
212 PAUL M. QUAY

word for this notion - and exactly iterable music and flows of sound is very
possibly partially responsible for the shift of interest to the historical and
temporal aspects of physics. For now one is becoming increasingly able to
think not merely regulatable objects in cognitive space, such static intelligible
forms as concepts, judgments, laws, theories, and the forms of reasoning but,
also, those patterns of temporal development and of primarily aural sequence
which seemed previously wholly evanescent.
E. Changes of the third class, by-passed till now, are the most fundamental
of all: those taking place in the physicist who, by changing himself, produces
the changes in the objective structures of data and theories and is at least the
proximate source for all the other types of change save the very first. The
relations between his own personal growth and each of these other classes
suggest an interesting rephrasing of our original question: How does an
individual physicist grow and develop in interaction with a continually
changing discipline? It can, evidently, be a strongly non-linear interaction in
several ways. For example, both Planck and Einstein each provoked a change
of the field drastic enough that each was no longer able to live or think
comfortably within its new framework. And each lived long enough for the
broader social effects of his impact on the field to react back upon his own
life - if nothing else, the atom-bomb would have been inconceivable without
both quantum theory and relativity.
Then, too, our consciousness of what we know changes as we become
more reflexively aware of the factors which condition and qualify our
knowledge, whether empirical or theoretical. Perhaps the most important
changes induced by relativity and by quantum mechanics were the changes in
our consciousness of what we know and how, making us intensely aware of
the always-prior suppositions of what some had thought to know without
presuppositions.
For this and similar reasons, I think one can discern historically a shifting
structure of epistemological development in the individual physicist which is
both illumined and complicated by the fact that the structure of temporal
interaction between the community of physicists and society at large is not
unlike that between the individual physicist and his professional community.
Fascinating though I find all this, rather than discuss it in somewhat decep-
tive generality, let me center solely on just a single, concrete example of
what I have in mind.
We are taught from our undergraduate days to understand and use
a number of logically contradictory theories. At first, I suppose, we use
not much differently from the way a carpenter uses hammer, saw, and
TEMPORALITY 213

chisel - as different tools, each for a different, well-defined task. Tisza's


logical justification for what is going on (Tisza, 1963, pp. 158-162) -
through subordination and dominance, with controlled inconsistency - is
sound. Yet as the years pass, we often have scant hesitation about using any
and all theories and models with which we are acquainted, with small concern
for the status of their interrelationships.
Eventually, however, the different theories come to seem like a collection
of lenticular interference ftlters. When trying to understand some problem in
depth, we switch, almost automatically, from one to another of these ftlters,
whose properties we have already analysed, seeking a different and helpful
view, through each, of the single reality we are looking at and worrying very
little about the mutual incompatibilities of the lenses themselves or the fact
that, if we seek to look through two at once, we see nothing. For, our formal
invariants, like such lenses, are not ordinarily what we attend to directly but
only the medium in which or through which we know the physical world.
And we need them all since, as Polanyi (1958, e.g., pp. 63-64,310-317)
never tired of insisting, anything real offers an infinity of aspects which are
not graspable from the original perspective through which it was first dis-
covered - a more sophisticated version of that much misused tale of the four
blind men and the elephant. A fuller investigation than is possible here would
show, I think, that this somewhat casual switching back and forth among our
theoretical 'lenses' is a relatively new development among physicists, Gibbs
himself feeling constrained apologetically to defend his modest use of the
method. In this way, then - and there are others - individual physicists'
epistemological orientations develop now in a manner significantly different
from formerly.

IV

Now, despite all these ways in which physics changes, we can say true things
about both physics and its changes. More importantly, truth is a necessary
component in each of those types of physics-activity. Objective truth, then, is
not generated by the art of physics as is foam on miles-deep of subjective
ocean. Rather, truth suffuses that art and directs it, at each level and in all
aspects, however limited and imperfect these contacts with reality may be
and however many our errors.
A. The knowledge that is most prized and which constitutes the reason-
for-being of the whole field is, of course, that which is most universal and
most necessary, hence, least subject to change. But the thrill of the univer-
214 PAUL M. QUAY

sally and necessarily true oUght not make us devaluate and forget particular
and contingent truths: It is true that I see, enjoy, and can be puzzled by the
varied colors of a laser's non-linear interaction with a crystal, or by the visual
or auditory manifestations. of oscillating chemical reactions. The man in the
lab knows that his Geiger-counters are painted blue, as well as knowing what
results he can expect from each of them. The theoretician understands his
mathematics and the varied, often disordered, contents of his physical pro-
blems and insights - and also that he uses A for the Helmholtz energy rather
thanF.
The entire activity of physics, knowing and searching for knowledge, is
governed and directed by the knowledge, particular as well as univerSal, which
one already has. It is present knowledge that provides us with problems and
with the verbal and mathematical languages in which to express them, that
makes us wish to learn, that suggests changes in what we had thought already
to know. It is our current understanding of the nature of our work that leads
us to interaction with the commonality of physicists and which gives the
ambient culture, often quite unperceived in its presence, one of its most
powerful grips on our minds.
The distinction between truths about matters of varying degrees of
contingency and particularity, on the one hand, and those of logical or
mathematical generality, on the other, is often confused with the related yet
quite different distinction between sensory and intellectual knowledge.
Sensory knowledge is always, indeed, of what is concrete and particular. But
I can know intellectually such particular, concrete, and contingent matters as
a free choice I have just made or the particularities of a present thought.
On the other hand, the intellectual knowledge that I have of an ache in my
back, of the taste of an orange, of the pleasure of warmth on a wintry night is
of the same basic sort as my knowledge of physics or of mathematics. I can-
not make you sense my pain, nor can I let you taste through me some
tropical fruit you have never known. Yet, in all which the mind can directly
grasp in such particular experiences you can share. And it is the precise
business of writers and poets to communicate to others such knowledge.
It is true that, to whatever degree knowledge is quantitative or mathe-
matically structured, to that degree it is easier to communicate to others: the
begetting of formally identical concepts is made much easier when the endless
complexity of concrete sense-experience has already been largely stripped
away, leaving almost uniquely defined structures. Yet even fme mathema-
ticians, maybe they especially, will resort to highly pictorial, intuitive, non-
rigorous language to generate insight into the unfamiliar areas of their
TEMPORALITY 215

thought. Rigor is the necessary and often painful mopping-up operation to


see that the territory acquired is empty of significant resistance, that the
insight is as true as it seemed to be, that nothing has been overlooked or
inadequately qualified.
B. ~ach of the five categories of temporal change we discussed earlier has
the peculiarity that it can be made to swallow up each of the others. Since
the time of the events to be described by physics is the real time in which
physics changes and physicists work, one can describe all categories, as I did
above, in terms of the simple temporal eventfulness of ordinary experience.
Yet this ordinary experience can be redescribed in the 'objective' time of
physics, an approach which led many in the last century to see the pro-
fessional growth of physicists, their history as a community of scholars, their
interaction with the environing culture, indeed, all of history as but the in-
calculably complex unfolding of the laws of mechanics for a system of as
many particles as the universe in fact possesses. History was but a solution of
an equation, the deducible consequence of the insights of physics. Even such
major revisions of physical theory as the introduction of relativity or
quantum mechanics become, on such a view, but the necessary results of
physical laws. Our current physics would be simply the way the real laws of
the world have forced our minds to think, whether there is any formal
correspondence between the laws our minds necessarily now hold and the real
ones or not.
After a while, historians returned the compliment and reduced physics to
an epiphenomenon of cultural growth, one way among many in which man
has viewed the universe and explained it to himself, a way certain in its turn
to be superseded by yet other, not related nor necessarily inferior, modes of
thought. Confirmation for such a view was found, of course, in the fact that
physics showed itself to be, even then to a keen eye, as mutable and tem-
porally variable as any other diSCipline. The realization of the particularity
of that one culture in which modern physics arose and the destructive and
corrosive effect it has tended to exert on other cultures, aided considerably
in establishing this approach.
Another example: All the above classes of temporal change can be seen as
but characteristic stages in the cognitive growth of an individual phYSicist.
Thus, in our ordinary experience, as we grow up we first know things as they
present themselves to us uncontrollably, endlessly diverse even in relation to
time (some change incessantly, others never) and unconnected save by
association or by often superficial analogies. One's self, too, is known in a
similar way, without reflection and as but one small part of the world. How-
216 PAUL M. QUAY

ever solid one's knowledge may be at this level, it is not seen as necessarily
true nor as universal in extent.
The scientific enterprise, however, undertakes the effort of rational and
objectively verifiable description and explanation, knowledge that is con-
ceived as that which any possible mind whatever would have to acknowledge,
given the same background. As a student, one's attention and interests center
almost exclusively upon the contents of lectures and texts, the results of
one's experiments, and the current state of the discipline. History of physics
seems a bore and, in any event, has nothing to offer as physics - at this level,
outdated theories are simply wrong.
But as one enters upon advanced research or study, the changeableness of
the field comes into focus, if for no other reason, because he hopes by his
own work to change it further. As he continues to grow, he becomes
increasingly aware of his own powers as a physicist and their ripening, also of
the effects of his field of activity upon the other areas of his life. Naturally
enough, this same process will make him more aware of other physiCists, of
his dependence upon them as well as theirs, to some extent, upon him. The
intersubjective aspects of his work, with all that they imply of the still deeper
interactions with the entire culture, also become stronger as he sees, for
example, that there are many ways of aiding the growth of physics, both as
knowledge and as art, which are not the doing of physics at all.
In similar fashion, all can be seen as changes in the community of physi-
cists and in the modes of communication with one's fellows. Putting the
matter a little differently, by taking a cross-section orthogonal to the above
approaches, one can, e.g., regard the changes in physics as an objective cogni-
tive structure not only as mere events in our temporal experience but as the
objective consequences of our present physical theory (or some potentially
future one) or as the subjective contributions of individual physicists' own
growth in understanding and delight in the world of their studies, or as the
results of communal interaction among professional men within a larger social
group or as the embodiment of over-all cultural shifts, much as Holton
suggests for special relativity.
This mutual swallOwing, one of another, is a well-known characteristic
of the cognitive disciplines. Each python swallows the python which is
swallowing him. Consider how the entirety of human intellectual activity
has been reduced to the workings of an absolute philosophical mind, or
to a will to power, or to economic struggle, or to sexual strivings, or to
socio-cultural evolution (with what strange but little considered implications
for the world that is known!). Each such position, though manifestly in-
TEMPORALITY 217

correct if taken by itself, points to one or other of innumerable aspects


of human nature and the world which, because quasi-universal, can seem all-
encompassing.
v
Our analysis thus far, however, is incomplete. We have spoken of the
structure of physics as a type of human activity entirely from its cognitive
aspect, as knowledge or the search for it. But we had started out to speak of
physics as a human endeavor; and this should remind us of the volitional and
other affective aspects which we have thus far ignored. At this point, I am
going deliberately to skip over the important but difficult question: Does it
make a difference (and if so, of what sort) to a person's success in physics
whether he is mature or not, is personally withdrawn from people or pro-
fessionally open to and critically interactive with others, culturally well
developed or retarded, morally good or bad? In brief, does the private or
subjective quality of the physicist-as-person have anything to do (if so, what)
with his ability to discover truth about the physical world?
But, in our present context, another question is certain to be raised at this
point which cannot be skipped: Whatever the influences of one's subjective
life or personal quality upon the origins of a given physical proposition, still,
does not the truth (or falsehood) of the proposition depend only on its con-
formity (or lack thereof) with reality and not at all upon the moral choices or
'lived' experience of its physicist-discoverer?
This is, in new form, but the old trap pointed out earlier. The truth of a
proposition does indeed depend only on its conformity with reality. But
propositions do not exist by themselves. If there is no living mind of which it
is the formal, cognitively invariant act, there is no proposition, only the
symbolic shell of one. Popper's great merit has been to drag that fallacy into
the open, even though defending it. But, in my judgment, there is no 'third
world'; there is some given human person and 'all-else'. This person knows
some of the 'all-else' by being intellectually conformed to it in greater or
lesser degree. No human mind, however, is capable of bringing itself into total
conformity with any object. Hence, there is no absolutely integral truth,
perfectly known, which could, therefore, be taken as totally transtemporal.
Since only this last could be utterly independent of the subjectivity of the
knower, questions concerning the personal aspects of his activity are never
totally irrelevant. Even to know what his proposition means requires
psychological activity and personal engagement.
This is not, however, a lapse into psychologism, which I entirely reject
218 PAUL M. QUAY

as a philosophically valid possibility. What we know is (in first intention) the


world. And our means of knowing it, our ideas and other cognitive invariants,
have their structure from the world, a structure which depends on the mind
only for its cOming-to-be within us and for its limitations. Yet to know what
we know, objectively and formally, requires self-knowledge and the use of
our affectivity, in particular our freedom, in the ways we shall see in detail in
Section VI. To describe our activity in knowing, a fortiori in seeking
knowledge, as purely cognitive would be just as inadequate as so to describe
any other of our human endeavors. For the knowing subject is also the
subject who is free.
Now, at one level, this is a commonplace. Every physicist recognizes that
choice is a major and essential ingredient in all that he does. He must be con-
stantly accepting, even if only tentatively, or rejecting data, ideas, arguments,
theories, themata of his own making or of others', forced to this, if for no
other reason, by the sheer number of such items and the limitations of his
memory. Data-gathering, also, depends crucially upon choices among possible
or plausible background material. No estimate can be made nor prediction
nor calculation without the choice of some consistent theoretical framework.
A key area of decision is that concerning whether a given problem is worth
an effort. Is it of sufficient interest to the physicist or his employer or
colleagues, in its own right or in its possible applications? Are the means now
available adequate, given a history of unsuccessful efforts by himself or
others? Is the process envisioned too tedious or time-consuming, given other
commitments or interesting topics for investigation? Does he himself have the
background to be able to attempt it? Are others already working on it or
about to?
Should one tackle the problem, once chosen, directly or using such
indirect means as deliberate over-simplifications or physical models? Should
he first make preliminary runs on ancillary aspects, such as calibration-curves
or subsidiary lemmata? Is this amount of data sufficient, accurate enough,
independent enough of merely contingent variables, sufficiently correlated
with the competing hypotheses under consideration to be able to aid a
decision between them? Should he work with full rigor and exactness from
the start or begin with plausible arguments and approximations? Are
mathematical analogies between physically quite different quantities to be
stressed or quietly ignored as irrelevant? Are the data accessible to him un-
biased with regard to the properties he is theoretically concerned with or
must they be retaken or reinterpreted? Will a restructuring or reformulation
of a theory make a real contribution or must one obtain new fonnulae and
make new predictions?
TEMPORALITY 219

Of greater importance for our present subject, however, is the fact that not
only does a physicist have an endless amount of decision-making to do in his
work with regard to each area of temporality but also his free choices
underlie and are causative of all these influences of time upon physics, albeit
not the only causative factors. His choices determine the configuration of his
own mind's grasp of physics. Thus, the way time appears in a theory is
evidently a matter partly of his ability but partly, and more importantly, of
his decision to have it so appear. Workers in mechanical statics did not
concern themselves with time at all. Those in classical thermodynamics were,
if not content, at least resigned to letting time have only the most shadowy
presence in their theories. But, as is now quite evident, these limitations
represented choices to make do with less in order to achieve at least some-
thing. They did not represent any renunciation of the deliberate and truly
communal choice to make time an explicit variable in any fundamental
physical theory.
Then there are the decisions to be made concerning major shifts in view-
point, as with Kulm's notion of the choice to be made between paradigms
when confronted by a potential 'scientific revolution', say, or Lakatos' choice
between competing research programmes. One's choices here, with those of
his fellows, determine whether, in fact, a 'revolution' takes place or not,
whether a given theoretical framework is adequate or not, which anomalies
are essentially unimportant, even if annoying, and which will require a careful
reworking of an entire theory.
More generally, each alteration we make in the relations between the com-
ponents of a complex cognitive invariant lies within our power to accept or
reject; so, for every judgment we make about theoretical or experimental
adequacy.
Further, the kinds of cognitive maturation in the life of the individual
physicist, one of which I discussed earlier, all depend in part upon free
choices. A physicist's development is not automatic. Only gradually does he
learn how to use his freedom constructively in any given context and then
move on to a further one. By the type of free activity in his field in which
he is already engaged, the limitations on the kind of knowledge involved
become known from within. It is not, say, that I find a problem I can't solve
with the methods at my disposal - that's a spur to further work. Rather, I
find questions of such sort that the class of possible answers, defined by the
kind of physics I have been doing until then, is ruled out as not even related
to the question. These limitations from within one's physics itself serve as
the motivation for each transition in turn.
In the first and simplest case: one's ordinary experience of movement,
220 PAUL M. QUAY

weights, machines is sensed to be inadequate, not because exhausted or


circumscribed by an imposed boundary but because the knowledge he now
desires should be of a different sort, should somehow give the real reasons
why machines move as they do, give solid ~ounds for predicting what any
given contrivance will do, let him know what things, in principle, can't be
done. Something similar is true for subsequent transitions, each calling for
a stronger exercise of freedom than the one preceding.
Less immediately connected with the content of physics, but in some ways
more important for its progress as a science are choices such as whether to
initiate some work in common with another researcher or team, to continue
such a collaboration, or to dissolve it. Should one continue seemingly fruitless
efforts, when one can see no way out of a hole, or is this rather the dark time
which is precursor to any real breakthrough and success?
But there seems no need to take the time in such a paper as this to spell
out the ways in which deliberate decisions by individual physicists modify and
move forward the history of their professional community' and affect its
orientations. Let the decisions which brought this Colloquium into existence
some twenty-five years ago suffice. Those decisions among physicists-and-
philosophers such as Frank and Bridgman, seconded or modified over the
years by those of others, have had no small influence on the ways physics has
grown in this country and the way it has understood itself.
How much the choices of the individual physiCists, singly or as a group,
affect an entire culture is equally obvious, though not always easy to under-
stand, from Galileo's choices concerning his work and its mode of publication
to the decisions of those who work for the achievement of fusion energy-
sources.
So, I think it is clear, even at the objective and the most elementary inter-
subjective levels, that physics can only be done by a free agent. As seen often
in students, to the extent that a person is lacking in self-confidence, is fearful
of errors, is indecisive and unwilling to take risks, lacks the ease of a mature
ability to choose responsibly and almost casually, to that extent his physics
will suffer.

VI

Since freedom appears as basic for all types of achievement in physics and
since it is the mover for all these temporal processes with which we are con-
cerned, it seems clear that we must examine more closely the relation
between knowledge and freedom_ The freedom of which we are conscious in
TEMPORALITY 221

our work is, at its most obvious level, an ability to choose among intellec-
tually possible alternatives, not all necessarily equally plausible or equally
likely to be right. There are, of course, other, more basic aspects also of our
freedom, essential for a complete discussion but into which we cannot enter
here. For example, free choice is also an act of self-determination. Or, there
is the problem of the ambivalence of our experience of freedom: we are
aware of our continuous potentiality for freely choosing, an awareness com-
mon to all who can know truth as such; yet we are aware of the full actuality
of our freedom only when we in fact choose well, in accord with truth.
What we do need to note here is that the endeavor which is physics is not a
discontinuous, aleatory jumping about of the mind. Freedom in the decisions
of which we have been speaking is not experienced by the physicist as mere
arbitrariness nor as a random event. If Einstein chose to reject the Copen-
hagen interpretation of quantum theory, it was not the result of his throwing
dice. Nor, though his reasons seemed to him powerful and persuasive, did he
regard them as apodictic save within a context which he recognized as already
freely chosen. Bohr, on the other side, and his major followers seem to have
adopted a similar stance.
Now, the most usual understanding at present of such freedom is that it is
a practical, even pragmatic, option or determination: "I don't know which is
better and, so, after doing my best to assess the likelihoods, I choose, since I
have to do something." Ultimately, however, this again reduces the freedom
of one's choice to arbitrariness: reason does as much as it can; the rest is
equivalent to the tossing of a suitably weighted coin. Yet this also fails to
correspond to our experience. Such situations do occur in the lives of all of
us, but we distinguish them easily from those deliberate choices of which we
have been speaking. Moreover, we remain uncomfortable in the former cases,
alert to see whether we should not revise our judgment and decide in another
sense.
A phYSicist, however, whatever role random events may also play, is
motivated and directed in his free choices by a deliberate desire to know,
truly and as fully as pOSSible, the universal and necessary principles governing
the properties and behavior of the most basic constituents of the material
universe, both in their simple interactions and in all degrees and modes of
aggregation. He is governed, even driven, by his determination to achieve an
objective knowledge and fully integrated explanation of the physical universe.
This deliberate desire and this determination, themselves the results of prior
and profound choices, 7 serve as global motivation for all his reasonable
choices (if one may consider him only as physicist for the moment). His
222 PAUL M. QUAY

daily, lower-level decisions, then, though free, are not unmotivated. In each
case, he sees some reason for each of the various alternatives, a reason which,
if reasonable from the viewpoint of physics, is that alternative's known or
conjectured relation to his goal of knowledge. And whether a good choice or
a poor one, in each case, the reason for it was its motive. Hence, critical
testing, attempted falsifications and the like make sense only insofar as they
can be shown to be apt means for leading us towards true knowledge.
Such phrases as 'seeking to know', 'desire to know', and 'determination to
achieve .. .knowledge' suggest that our freedom is engaged somehow in our
knowing even as our current knowledge offers motives for our choices. What I
want to argue now is a bit more precise: that knowledge and freedom are
dialectically related in this sense, that there can be no intellectual knowledge
without the presence of freedom as its ground of possibility just as there can
be no freedom, deliberate and non-arbitrary, except on the basis of intellec-
tual knowledge of possible goods. Having said a few things about the latter
point, I would like to expand a bit on the more controversial prior one.
A. If it is to make sense to say that one is able to know anything whatso-
ever, the first epistemological requisite is that what is known have 'entered'
as it is, in its otherness, the mind of the knower. From early Greek
philosophy on, men have been concerned with and have much debated the
sort of presence something has within the mind when known. 8
Yet, whatever be the explanations offered for knowledge, to know some
real state-of-affairs at all, I must receive it, hold it, permit it to be - in the
mental mode of presence - as it is, without distorting it or re-forming it by
subjecting it to my own prior conceptualizations or forcing it into my already
extant, mental representation of the world. If, in knowing, I permit my own
prior knowledge, my subconscious predilections, or my freedom to make the
'known' conform to me or to particular aspects of my mind as it is presently
constit1:lted, then, to that precise degree, I have obliterated the other which I
think to know, and wind up knowing only myself or some aspects of my own
mind or, at best, some badly refracted 'image' concerning which I know
neither the manner of its refraction nor even that it has been refracted.
Yet, this openness of the mind to the physical world (or, for that matter,
to the current content of physics) is not a simple passivity, as the ancients
thought at least sensory knowledge to be, a receiving of the formal imprint of
the object as if it were a seal on wax. For we now know that the senses
receive actively, processing all input as it is being received, constructing,
repressing, discriminating, so that, e.g., we learn quickly to see objects and
not mere colorations. One major difference between the sensory receiving and
TEMPORALITY 223

the intellectual lies in the deterministic, albeit spontaneous, character of the


former as over against the free activity of the latter. There is, of course, much
spontaneous, non-free processing of things intellectually also, a fact which has
a major role in the psychology of discovery. In both cases, the automatic
operations arise from an interaction between the structures caused by what is
already known and what is now entering one's cognitive field. But under most
circumstances, the person remains free to accept or reject or modify the
cognitive form seeking entrance.
Further, the physicist seeks to interiorize everything that he knows, that
is, to bring it into articulate and reflexive consciousness. This interiorization
often involves an active wrestling with prior conceptualizations, which the
new notion must at least displace, if not destroy, so that a certain restruc-
turing of one's thought-patterns is required. This almost automatic response
witnesses to the depth of our conviction that all that is true must cohere in a
single unity.
In this two-fold process of opening and interiorizing, a person touches
reality only by consenting to be changed himself, by consenting that the
other-than-himself remain such in his knOwing it. This letting-be while fully
attending-to, this respect for what is other, as such, is already an adumbration
of love, both as contemplating and possessing. Like love, also, it is a going out
of the self, a leaving of one's prior framework of thought, expectation, and
hopes, in dependency on that other.
Good as all this is, it does not always appear to us as good, at least not as
unmixed good. Since it is I who am structured by my knowing, it is I who am
'worked over' by new knowledge, who am threatened by the ever recurring
possibility of having to rearrange all that I know or thought to know - still
more, to rearrange my life, even myself, by letting in some new and seemingly
alien element which does not fit with what is already present.
Simple knowledge of something, then, requires a pure attention to the
being of the thing as other. But such pure attention is a matter of my choice,
both as attention (for I may divert my attention elsewhere) and as pure (for
I may warp the mode of its being in me by the way I receive it and adapt it
to myself). I need not, then, give this pure attention, save (possibly) in that
most basic intellectual knowledge of the world out of which, through the
effort of articulation, the so-called first principles are crystallized.
B. Yet a simple knowing of something is incomplete. It needs articulation
and the assertion, at least to myself, of what I think to know, if there is to be
question of truth. Here especially, my freedom is operative.
Let me begin with the assumption that the mind is capable, by its nature,
224 PAUL M. QUAY

of truth. I do not mean that the mind can resolve any puzzle posed it or find
out all that is. The truth may well be: "I don't know the answer" or, often,
"I am inclined to say so but cannot" since the evidence is inconclusive. What
I do mean is that, when the mind is operating properly, that is, in response to
the person's love of truth,9 it would invariably be able to know the kind or
degree of its own conformity (or lack thereof) to the reality which it is con-
sidering. This knowledge of conformity is a reflex knowledge in which,
following upon my assertion, I perceive this as certain or plausible or
erroneous. Adverting to my direct experience, I become aware of grasping, in
the act of knowing, the fact that I do know or, on other occasions, that my
mental activity cannot yet be giving me more, on a specified point, than
opinion or surmise.
I take it also as evident that each of us has, in fact, erred. If we were so
constituted that we were not merely necessitated to our judgments but were
necessitated always to judge truly, our problem would be somewhat different.
But few physicists, I think, are deeply concerned over that difficulty.
We are equally aware that we are not necessitated to error as such. We do
arrive at some truths, even if they be no more than truths of experience or of
logical consistency. Further, without having known truth, we could scarcely
be conceived to have constructed such a notion, still less that of error. Nor is
it conceivable what could establish a bias towards formal error in us, error
being indeterminate until the truth is known. To exclude truth always, one
would already have to know it - unless one wishes to worry himself with
Descartes' "evil genius". It is true that there are infinitely many more ways
of being wrong than of being right. Error might, therefore, be more likely,
statistically speaking. But, a reasonable approximation to the truth would
have to recur every so often or some profound reason for such non-ergodic
behavior would be required. If our judgments were randomly true or false,
natural selection would soon take care that those who survived would be
biased in favor of truth.
Now, if we are necessitated neither to true judgments nor to false ones, we
may yet be necessitated to our judgments but independently of their truth
or falsehood. At least one other variable or degree of freedom is then called
for, e.g., anger, desire, fear. For, if simply necessitated to our judgments,
assert them we must, whether true or false. But a judgment, by its internal
structure, is an enunciation of what one thinks to perceive as true. If, then,
we are determined so as to have at times to perceive (and declare) both true
and false propositions alike to be true and at other times to judge both kinds
alike to be false, as determined by the 'hidden variables', we clearly have no
TEMPORALITY 225

genuine power of discrimination between the true and the false, sensitive
though our minds would be to the 'hidden variables'. We should be like
machines, preprogrammed to respond in a certain way.
A little hand-computer can give correct 'answers' far oftener than I to
most of the problems I use it for - but it cannot do otherwise. And if there is
a breakdown or flaw in the circuitry, it may offer as many wrong 'answers' as
formerly it offered right ones. The 'answers', of course, are not answers at all
but certain physical behaviors of the machine that have been designed by some-
one to be isomorphic with a given arithmetic process and conclusion, which
isomorphism enables the human user to interpret every resultant behavior in
terms of the numbers which correspond to the output-display - as long as the
machine is working properly. An electronic turtle can be programmed in a
way that turns out to be, were it capable of reproduction, evolutionarily
advantageous, i.e., which keeps it from being destroyed by the more ordinary
hazards of its environment. But neither is this what we mean by knowing the
truth. And surely, those who have developed the theory of evolution of
species do not mean it to be true only in that sense.
It is only by knowing and by being free to evaluate and put aside the
possible intrusions of error that I can know that I am knowing truly, that I
am intellectually conformed to some reality and so am able to make a
formally true assertion of what that conformity embodies. The power to
know truth requires, in man, the power to avoid error. In the very knowing
that, in my judgments, I am not determined to the truth as truth, I know that
the knowledge of truth requires that my freedom act to bring about that deter-
mination, by setting aside all other determinations. If I have no such freedom,
I can never know that what I am necessitated to consider as true is such.
Scientific knowledge, then, is had only through personal (Polanyi, 1958),
and free activity. Methods, in the strong sense of processes guaranteed to
reach truth if rightly employed, can be isomorphically imaged by machines
that can operate correctly, i.e., in accord with a predetermined mode or
pattern. But, for exactly the reasons mentioned. above with regard to
machines, truth cannot be had from such a scheme. 1o The notion of method,
in the strong sense, therefore, is internally inconsistent. Such a method would
seek to guarantee that we arrive at intellectual knowledge automatically. Yet
this guarantee, profferred without regard to the nature of the unknown,
could be sought only in a proper ordering of our own internal operations,
whereas knowledge and truth, as the free reception of what is other than
ourselves, depend upon what is, at least in part, not internal to us as
subjects.
226 PAUL M. QUAY

VII

Let us return to our initial problem. Most of us would concede at once, I


suppose, that the art of physics is essential to the science of physics. This
latter, however, many would restrict to those reproducible data and trans-
parently rethinkable arguments which constitute the objective knowledge of
the physicist. But, precisely because free, because not produced by coercive
argumentation, the decisions of the physicist's art are declared to be
irrational, even though inescapably so. Physics is then seen historically as
developing, in some inexplicable fashion, by largely irrational processes into
the most rigorously exact manner of dealing with the world yet contrived! 11
So, the key question is: does the non-rational element in physics (in the
physicist, acting as physicist) destroy or damage the rational?
The distinction we need is that between the irrational and the non-
rational. Free choice is not, taken formally, an act of reason or of under-
standing. It is, therefore, non-rational. But a choice is irrational only if some-
how contrary to reason, that is, without an intellectual perception of an
adequate goal as one's motive. But, given the goal of physics, more or less as
earlier described, then all the non-cognitive or non-rational efforts and activi-
ties of the physicist, alone or in community, have a properly rational aspect
insofar as he sees or conjectures them to be helpful to reaching his goal. For,
the rationality of any activity lies in its intellectually perceived suitability for
achieVing one's present purpose without hurt to one's more nearly ultimate
goals.
The goal of the phYSicist is, of course, a goal only because he chooses it,
desires it, seeks it - even as a problem exists only for one who is bothered by
it, senses the need to resolve it, becomes involved with it. Just as a purely
intellectual perception of a conundrum results in nothing more than the judg-
ment: 'This is a conundrum', so, a purely intellectual seeing of the good of
knowledge and of a quantitative understanding of the physical universe
produces, of itself, no more than the judgment, 'Such activity is good; and
someone might rightly make it his goal.' Hence, again, there is need for love
of truth. It is in the fact that truth which is empirically grounded and
theoretically understood and framed with logical consistency is a good for us
that our puzzle finds its solution.
Knowledge of physics, as a good, is eminently desirable. Our desires for it
are not themselves cognitive nor do they result from mere logical argument,
no matter how correct. Yet it is only in virtue of these desires that true
knowledge will be had, that any rational search for physical truth can be
TEMPORALITY 227

pursued. The desires in question can come from all levels of our being:
intellectual perception of the good, fear of a universe still profoundly
unknown and hence uncontrollable, desire for an earthly immortality through
one's fame among men, the desire of a well-paying job in an activity one
enjoys, the furtherance of some related activity such as weaponry or
engineering, and endlessly many others. Whether and in what way each of
these, save the first, may be a rational goal for doing physics will depend
upon the rationality of one's prior goals and the way each of these desires
truly accords with those greater goods.
The dialectic between knowledge and freedom, between truth and love, in
every concrete act of knowing or of choosing, also provides the foundation
for responses to questions about moral values in physicists' decisions, the role
of scientific knowledge in moral judgments, whether 'simplicity' or 'beauty'
are proper criteria for the assessing of a physical theory, and the like.
For example, ought the beauty of a physical theory, with all its overtQnes
of subjective pleasure or enjoyment, be allowed to weigh as evidence for its
truth? Of course, if "beauty is truth and truth, beauty", we would be wrong to
disjoin them. Nor is there any rational ground why a physicist should not be
drawn to a theory by its beauty. The difficulty is that the intensity of our
response to beauty is so much greater than our response to truth when not
yet perceived as beautiful that even a vestige or trace of beauty in a theory at
least largely incorrect can blind us and make us take it as wholly true. More
commonly, we do as Kepler and let the beauty of one aspect (e.g., the mathe-
matical) substitute for the beauty that we should be seeking (e.g., the
physical).
The error of the historicists generally, I think, is to see free choice as
arbitrary, unmotivated, essentially random. We have already indicated that
freedom is not arbitrariness and that whatever is properly chosen has adequate
motivation in our knowledge of each alternative. Just as it is in free choice
that truth is grasped and asit is only in freedom that intellectual knowledge
is possible, so only through one's grasp of truth is there proper exercise of
choice and only in knowledge can freedom be integral.
But, one may insist: "Ought a physicist's awareness of himself, his feelings,
his freedom, his interpersonal relations to have any least significance for our
evaluation of the truth of his physical theories or his data?" Perhaps not. But
notice that, already, one does not dare say, "Ought those personal aspects
have any least significance for the advancement of physics?" Evaluation,
corroboration, and the like are processes which derive from and issue in our
free decisions to accept, reject, or weigh further his theories or data (though
228 PAUL M. QUAY

the 'he' of today may be merely the 'I' of yesterday). Note, too, that while I
may rightly exclude from my considerations of validity his feelings of delight
in the beauty of his theory, in the elegance of his experimental design, in his
constructing of a truly new idea, I do not ordinarily exclude my feelings con-
cerning these same matters as simply irrelevant.
A somewhat adolescent passion for demonstrable certitude about every-
thing and for total objectivity has tended to neglect all aspects of physics save
that of evaluation of validity. Such a passion has no demonstrable validity
itself but represents a subjective demand, imposed a priori upon the world,
that it conform to the essentially incomplete mental set which the young
person has already acquired. We require no better example of the need for
freedom in finding truth. Further, validation, like logic, is in all senses an
'after-thought'. Each is essential in our checking out the degree to which our
minds have made contact with reality and in resolving the inadequacies and
confusions of our ideas. But logic is only very weakly constructive, having no
categories for genuine insights poorly expressed; and validation has no way
to confirm or refute confused, half-articulated, only partially conceptualized
judgments. If we did not know, and know both truth and error, we would be
quite unconcerned with validation or corroboration or falsification - or any
of our scientific labors.
To conclude, then, many have argued to an. essentially non-rational
element in starting or carrying through a 'scientific revolution' or otherwise
changing the current cognitive structures of physics. I have argued that there
is an essentially non-rational element in constructing a syllogism - but that
there need be nothing illogical or irrational in either case. The non-rational
element of freedom does not injure rationality; instead, it grounds its possi-
bility, stimulates and furthers its action, and provides its principal protection
from irrationality.

Loyola University, Chicago

NOTES

* This paper was written for the Boston Conoquium for the Philosophy of Science, and
was presented in a slightly abbreviated form at Boston University on 3 April, 1979.
1 Denying truth to physics is, of course, an abdication of the philosophical effort, not a
solution to the problem. So, for example, Laudan (1977, pp. 106 -114 ) attempts to give
physics a worth in proportion to its historically ascertainable 'problem-solving' abilities.
But if we could not get truth from physics, we assuredly would not be able -to obtain
TEMPORALITY 229

it from history. Yet the actual problem-solving ability of any tradition in physics can
only be known if history gives us detailed truth about what physics has done. If truth is
wholly to be rejected as a possibility (ibid., pp. 123-126), Laudan's criterion based
on problem-so1ving ability is useless, since he could not assert truly that this approach
has any value. Apart from that self-contradiction, his approach induces an inftnite
regress, requiring that we ftrst choose our theory or tradition of history on the basis of
its ability to solve problems before we can make use of it to fmd out about the problem-
solving ability of physics. But, whatever criterion we use to judge problem-solving ability
of theories of history must itself be validated by its ability to solve problems better than
some other proposed criterion. This must go on ad infinitum, into more and more
obscure domains, or else terminate in something we can accept as true.
2 I do not mean that every physical theory deals with time. Various types of purely
static theories exist, e.g., mechanical statics, electrostatics and magnetostatics, crystallo-
graphy; and others choose to deal only with one or other aspect of time, e.g., the simple
sequentiality of thermostatics.
3 The problem is, at root, rather the historian's borrowing of the logicist's problem:
how can two contradictory theories be simultaneously true, as physicists seem com-
monly to assume? Under the impulse to seek certitude through rigorous logic (which
provides, at its best, a general type of information-preserving transformations but is
never generative of truth), the notion of partial truth seems to have been lost, as also the
recognition of the ordinary use of propositions which always involves incomplete formu-
lation (for, the speaker says more by his words than he intends while also he intends
more than he can say). Laudan gives an excellent discussion, in terms of research tradi-
tions, of some of the elements involved here (Laudan, 1977, pp. 70-106) though he
fails to see their import for this problem.
4 Even physics-as-Idea would not truly be such, due to its connections not only with
mathematics and logic but with biological and all other disciplines. The only totally
transtemporal knowledge could be the inftnite and nontemporal knowledge proper to God.
S It is important, in view o£ Laudan's opinion to the contrary (1977, pp. 21, 38, and
106-120), to note that our problems (or what we personally choose to regard as such)
and our preferences for a 'problem-solving tradition' are not, at base, competitive. Many
of the most interesting problems are recognizable in themselves and no competitors
exist: e.g., the problem of finding a general, time-dependent thermodynamics.
6 It is chiefly the neglect of this deductive structure in the process leading to discovery
- for discovery itself is not properly. a process but an event, not existing at all till it has
been achieved - that weakens Blackwell's otherwise very useful analyses of scientific
discovery (Blackwell, 1976 and 1981) into a sort of psychologism.
7 Since not everyone gives his life to such efforts, it follows that the physicist's goal, as
a practical object-to-be-attained, is the result of prior choice. This prior choice itself is
made in virtue of goals prior to it; and so on, for each choice of goals, until the ultimate
goal is reached, about which, I think, we have no freedom: the complete good, which
contains all truth as well as all else necessary wholly to sate our longing for the full and
balanced activity of all our powers, with and through and for others, a good concretely
attainable, I would argue, only in God.
8 That it is, in some manner, there, is patent - if I know it, I am different through my
knowing than did I not know. This modification of me by means of my knowing is a
presence of that thing, directly or indirectly, within me and to me, as well as an activity
230 PAUL M. QUAY

of my own. That the known is not present in the same mode in me and in the world is
even more evident. If nothing else, a tree or a house just won't fit in where I know
myself to be. Further, I can forget the tree or house entirely; it remains there for other
o~ervers - and for me, should I advert to it again.
9 The optimal activity of the physicist's mind occurs when its proper goal is most loved
and best loved, i.e., when the physical world is regarded with that attentive and reverent
fascination which respects fully the particular nature of that world in the very seeking to
satisfy one's love with deeper knowledge. Feelings, emotions, and freedom, in brief, our
affectivity, can and should integrally and harmoniously aid in pursuit of that truth. The
affective component may be less violent than in other kinds of love but need not be less
f oowerful.
As pointed out earlier, logic and some portions of mathematics are not concerned
with attaining knowledge or truth in the fust place but with rmding those transforma-
tions which leave invariant such knowledge or truth as has been expressly formulated.
11 We have already seen one essential link between physics-as-search and physics-as-
knowledge in the deductively structured movement of the mind from its most primi-
tive perceptions of truth, by means of estimation based on particular and contingent
truths, to the great structures of theoretical physics (cf. Section III, B, above). Though
this link is insufficient by itself, yet no adequate response to our question can be had
without it.

REFERENCES

Blackwell, Richard J. 1976. 'Scientific Discovery and the Laws of Logic.' The New
Scholasticism 50 333-344.
Blackwell, Richard J. 1981. 'The Rationality of Scientific Discovery.' In Wissenschaftliche
und ausserwissenschaftliche Rationalitiit: Referate und Texte des 4. Internationalen
Humanistischen Symposiums, 1978. pp. 189-207. Athens: Griechische Humanisti-
sche Gesellschaft.
Holton, Gerald. 1973. Thematic Origins of Scientific Thought: Kepler to Einstein
(Cambridge, Mass.: Harvard University Press).
Lauden, Larry. 1977. Progress and Its Problems: Toward a Theory of Scientific Growth
(Berkeley: University of California Press).
Polanyi, Michael. 1958. Personal Knowledge: Towards a Post-Critical Philosophy
(Chicago: University of Chicago Press).
Popper, Karl R. 1972. Objective Knowledge: An Evolutionary Approach (Oxford:
Clarendon Press).
Quay, Paul M., S.J. 1974a. 'A Distinction in Search of a Difference: The Psycho-Social
Distinction Between Science and Theology.' The Modem Schoolman 51 345-359.
Quay, Paul M., S.J. 1974b. 'Progress as a Demarcation Criterion for the Sciences.'
Philosophy of Science 41154-170.
Quay, Paul M., S.J. 1975. 'The Estimative Functions of Physical Theory.' Studies in
History and Philosophy of Science 6 125-157.
Tisza, Laszlo. 1963. 'The Conceptual Structure of Physics.' Reviews of Modem Physics
35151-185.
S. S. SCHWEBER

COMMENTARY ON 'TEMPORALITY AND THE STRUCTURE


OF PHYSICS AS HUMAN ENDEAVOR'

At the outset, let me say that I found Father Quay's paper a stimulating, sen-
sitive analysis of physics as a human endeavor. The tapestry that he has
woven includes many of the strands that I believe must be included in any
such undertaking. I particularly welcome what I would can the Jamesian
element of his analysis; he has insisted that to any sociological analysis of
physics must be added a description of the varieties of experiences of the
scientists themselves ("how it feels from the inside") and that to understand
the dynamics of the growth of physics, one must appreciate the function of
this 'feel' in propelling the science forward.·
Father QUay's stress on the temporal aspect of physics - and by extra-
polation, all knowledge - is surely right. In this analysis of temporal rela-
tionships which characterize physics - and by extrapolation, all knowledge
- he has set up various temporal classes or categories:
- In the first, it is the objects which a theory is set up to describe which
change in time.
- In the second class, the theory itself is changed in time.
- In the third, the theorizer (whether as formulator or student) is changed.
- In the fourth, the community which validates or uses the theory is
changed.
- In the last, the notion of what a theory is and what is expected from it
also varies.
It is the merit of his structural analysis - and its deficiency - that it
applies equally well to most forms of human activities in their search for
'objective' knowledge. There is undoubtedly a commonality to anthropo-
logical knowledge, chemical knowledge, mathematical knowledge, precisely
because they are the end products of human activities. But it is also the
characteristic differences of these activities that we wish to discern. We want
to understand the difference between the criteria of what is considered
interesting and deep in anthropology, in phYSiCS, in mathematics, etc. Again,
Father Quay would undoubtedly be right in asserting that history - the
evolution of the discipline, of the community of practictioners, etc. - has
played an important role in attaching meaning to the questions which are
asked and to the answers which are accepted by the community.
231
R. S. Cohen and M W. Wartofsky (eds.), Physical Sciences and History ofPhysics, 231-238.
© 1984 by D. Reidel Publishing Company.
232 S. S. SCHWEBER

But I suspect that Father Quay would not accept history as the only
determinant. His own stress on the transtemporal, formal stability of some of
our theories - and the role of such 'stabilized' theories in erecting the skele-
tons and musculature of our minds - suggests that he would accept the thesis
that in our searches, we have perceived certain ahistorical elements. And I
remind you that Nietzsche indicated "only that which has no history is
definable." 1
To obtain a better understanding of our knowledge of the world, of our
relation to it, and of the differences in kinds of knowledge, I believe it is im-
perative to analyze more carefully 'the objects' our knowledge deals with, the
theories that are set up to describe them, and more particularly, our relation
to them.
Richard Burian and I have attempted to formulate a model where these
'objects', the compositional 'units', the elements which are to be used in
explaining the phenomena in the universe of interest are critically analyzed.
We allow these 'units' to interact both energetically and informationally. The
choice of compositional units will depend on the environment, on· the
phenomena to be explained and the 'fineness' of the investigation being
pursued. The units together with the environment form the system to be
described. These notions are made more precise by illustrative examples.
Atoms, ions, molecules, protons, electrons, mesons are some of the structural
units used in physics and chemistry. In the usual situations encountered in
those fields, the interactions are (adequately) described as exchange of
energy, momentum, charge, quantum numbers (our label 'energetic' is
heUristic).
A program for a computer represents the structural unit which is the proto-
type for informational interactions. What is characteristic of this kind of
interaction is the transfer of information (Le., instruction to perform speci-
fied operations). Biology provides the best examples of structural units which
interact both energetically and informationally. At different levels of descrip-
tion, DNA molecules, genes, cells, organs, organism and species performing
distinctive functions in an ecological community are obvious examples.
Although it usually is the case in the physical sciences that the structural
units, when isolated, possess relatively simple characterizations, this need not
be the case. The requirement of simplicity is usually related to a theoretical
scheme such that the properties of the isolated structural units can be used as
a starting point for deriving properties of the composite system.
Our structural analysis implies that there is an immediate requirement
for any useful description of a system undergoing a series of changes: the
COMMENTARY ON 'TEMPORALITY .. .' 233

description must focus on those aspects that are at least initially relatively
unchanging, i.e., durable and stable. It is this aspect of stability which is the
most important characteristic that we shall demand of our structural units.
The description of systems in terms of (relatively) stable compositional
units allows a unified approach to physical, biological and social systems. But
more is wanted than a possible superficial characterization of such systems.
One wishes to construct theories (programs) which are able to describe,
predict and/or facilitate control of the processes the systems undergo. The
character of the theories will, however, depend drastically on the character
of the structural units which make up the system.
At one end of the spectrum in the description of social systems, the
structural units may change so rapidly on the time scale relevant for the
phenomenon that in effect one does not have 'stable' structural units. In the
event that the structural units change, but only slowly over the time scale of
change of the system as a whole, one speaks of the role of 'history' on the
structural units. In other cases, the number of structural units that have to be
considered to make the system approximately isolated may be so large that
no effective computational method may be available, even were one able to
write down a program for the evolution of the system. What is characteristic
of an interesting description of social systems, and of organisms at the
population-biology level, is that one is dealing with structural units which
have enough stability and which are sufficiently 'identical' to satisfy some well
defined criterion for class membership. Stated differently, the structural units
in these systems are all non-identical, and the theories must reflect this fact.
At the other end of the spectrum, the description of physical systems at
the microscopic level employs structural units which have been shown empiri-
cally to be strictly identical in kind: all protons, all electrons are identical, all
isolated hydrogen atoms in their ground state are identical, etc. It is precisely
this strict identity and the stability of the structural elements which makes
possible the fully objective character of the knowledge of physical systems.
It is precisely the stability and identity of the structural elements and their
ahistorical character which allow the constant repeatability of the processes
described by the theory, even though they are embedded in a changing
universe. The objective character of macroscopic physics (where laws can be
derived even though no two physical objects are exactly alike) still stems
from the identical character of the elementary entities making up the macro-
scopic entities. The details of the immense number of microstates compatible
with macrovariables average out and are thus irrelevant. The averaging
procedure is justified from the identity of the constituents. Conversely, it is
234 S. S. SCHWEBER

the lack of identity and the instability of the structural elements used in the
description of social phenomena which makes the programs there subjective,
that is, constantly in need of revision and adjustment, even in the description
of the units of interaction. This difference lies at the root of the distinction
which people make in their system of beliefs when they differentiate between
beliefs: "one relating to a world of objects, facts and concrete events, one to
a system of values, obligations, conventions and institutional categories." 2
We see the fundamental issues of theory building turning on the choice of
units. In actual fact, theories are neither built from successful predictions nor
overthrown by predictive failures (even by a considerable accumulation of
predictive failures). They are themselves fundamentally stable to such pertur-
bation as long as the structural units they refer to can continue to be treated
as the basis for both energetic and informational exchange. Predictive failure
can normally be overcome by rearrangement of units, a process often
extremely complex in detail, but uninteresting in principle. In the long run, it
is not falsification and predictive failure which overturn a theory, but the
location of 'better' units and a theory appropriate to them.
We accept the Bohme et al. 3 reformulation of the Kuhnian thesis of the
development of scientific disciplines. With them, we shall consider three
phases:
Phase 1: The pre-paradigmatic;
Phase 2: The paradigmatic; and
Phase 3: The post-paradigmatic.
The paradigmatic and post-paradigmatic have been considered together by
Kuhn as "normal science."
The pre-paradigmatic stage of scientific discipline is that period of its
evolution prior to the emergence of stable structural units. It similarly lacks a
generally accepted theoretical framework by means of which to structure its
investigations. In the words of Bohme et al., "its main goal is discovery not
explanation." This phase is thus characterized by the fact that empirical,
descriptive and taxonomic strategies are predominant. Historical examples
of the pre-paradigmatic phase are electricity before Franklin and chemistry
prior to Boyle.
The maturation of a scientific discipline witnesses the emergence of
structural units (though not necessarily stable ones) and corresponding
theoretical structures which also serve as guidelines for the research. Kuhn has
described this process as the emergence of a paradigm. Lakatos' characteriza-
tion of this phase as the emergence of a research program is probably a more
apt deSCription.
COMMENTARY ON 'TEMPORALITY .. .' 235

This second phase of the Bohme el al. model is characterized as that stage
in which 'fundamental theories' are formulated - theories which are not
completely superseded by 'revolutions'.
We would characterize this phase as the one in which stable structural
units have been apprehended, eventually resulting in stable fundamental
theories. Research during this phase is guided by the search for an adequate
description of the stable structural units and for an adequate inferential
apparatus (or some other technique) by means of which to exploit the
descri{?tion of a situation in terms of the structural units in explaining the
phenomena of interest. In short, the search for an adequate theory based on
the structural units already identified (if only partially) becomes, in the
terminology of Bohme et al., the "regulative" governing research. The guide-
line of scientific advance in this phase is, thus, not "the theory itself' as
Bohme et al. would have it, but the adequacy of the theory as an explanation
of the phenomena given that it treats the structural units in question as the
basis of explanations. This "regulative" belongs to internal history of science
and (as Bohme et al. argue), it is the dominating regulative in phase 2.
Classical examples are the elaboration of the 'central dogma' following
the discovery by Watson and Crick of the structure of DNA and the develop-
ment of quantum mechanics from 1925 to 1929.
The second phase can come to completion without arriving at a degene-
rating problem shift (i.e., without anomalies producing a crisis in the dis-
cipline) provided there exists a 'fundamental theory'.
Heisenberg's characterization of a theory as a fundamental theory of a dis-
cipline, i.e., as a "closed theory," is as follows:
(a) it is not improved by minor modification;
(b) it is not invalidated under revolutionary change (example: classical
mechanics and relativity, or quantum mechanics for that matter);
(c) it is valid for a certain range of phenomena and highly corroborated in
that range; and
(d) it is logically consistent to the extent it is axiomatically formulated.
When a stable fundamental theory has been achieved, it analyzes a range of
phenomena in terms of some set of underlying structural units. The stability
of the theory amounts to the fact that (1) the theoretical characterization of
the structural units has been developed in a tightly coherent manner, so that
minor modifications bring no gain with them, but require major adjustments
in the theoretical system, (2) that a range of phenomena of interest has been
successfully characterized in terms of these structural units, thereby (3)
constituting one or more domains of stable application of the theoretical
236 S. S. SCHWEBER

apparatus for description, explanation and control. Although it is hard to put


the point adequately in a formal or semantical way,4 the theoretical
characterization of the structural units must be doing real work in charac-
terizing the phenomena of concern for there to be a genuine domain of stable
application of a fundamental theory. (Thus, the Greek atomic theories did
not constitute stable fundamental theories in our sense since, unlike
nineteenth-century atomic theories, there was no range of phenomena which
was explained in terms of the independently formulated and stably charac-
terized details of atomic structure.)
The achievement of a fundamental theory is marked by the conjlu;nce of
a theoretical characterization of certain structural units and the actual
operation of stable units, picked out by the characterization, in some
domain(s) of application of the theory. This confluence justifies the claim
that the problem of explaining the phenomena in those domains has been
solved 'in principle'. But the gap between 'in prinCiple' and 'in practice' must
not be underestimated.
I have presented these notions because they clarify the time scales and the
dynamics of change of the various classes in Father Quay's analysis and also
because I believe it is part and parcel of Father Quay's enterprise to get a
better handle on what differentiates knowledge in physics from knowledge in
mathematics, in biology, in anthropology, etc. The case of physics and mathe-
matics is particularly interesting because it sheds further light on our relation
to the cognitive structures - our theories - and to the entities these theories
deal with. Whether you are a platonist or a formalist as a mathematician has
its parallel in whether you are a realist or not as a physicist. The presupposi-
tion of the nineteenth-century mathematician that mathematics must be.
anchored on an absolutely certain and reliable foundation paralleled that
century's physicist's commitment to reductionism and mechanical explana-
tions. In some sense, the case of mathematics is easier. 5 Mathematical objects
are created or invented by humans, but they are not created arbitrarily; they
arise from activity with already existing mathematical objects, and from the
needs of science and daily life. Once created, these mathematical entities
have well-determined properties which they possess independently of our
knowledge of them. Moreover, mathematicians mayor may not be able to
discover these properties. Mathematics is part of our non-material culture, a
reality that is inner from the point of view of society as a whole, but outer
from the viewpoint of each individual member of society. That this reality
has a stability and permanence is a social fact, a consequence of our evolu-
tionary and biological inheritance, somewhat akin to the language we speak.
COMMENTARY ON 'TEMPORALITY .. .' 237

Some facets of physics are more difficult to analyze and others easier,
because the 'real' world puts more constraints on physics than on mathe-
matics. It requires the identification and the analysis of objects which we
believe are part of the non-human outer world and the matching of the
observed properties of these objects to structures which we have created
to describe them in as simple, economical and utilitarian a fashion as
possible.
Incidentally, this enterprise could terminate. It could be that the search
for elementary constituents yields nothing new; or that the match between
theoretical entities and elementary structural units results in a cycle of Men-
deleev-like tables for the structural units and that these 'tables' are verified to
ever higher energies. Or society could decide not to invest further resources
into that endeavor.
In any case, as Father Quay has beautifully indicated, the endeavor and
enterprise is temporal and historical. Yet there seem to be transtemporal
elements - glimmers of truth.
Father Quay wants to find a basis for the achievements in physics in the
relation between knowledge and freedom. In a way that I cannot express
very clearly, the parallel seems to me more analogous to the problem of
linguistics: to understand how it is that languages have mysterious com-
plicated possibilities unknown to the speakers of the language. There is a
not understood deep connection between our biological make-up and our
culrure.
In fact, Schrodinger alluded to this dilemma in his Epilogue on 'Free
Will and Determinism' in his seminal book What is Life? 6 He there asks the
question, how is it that "I ... am the person, if any, who controls the
'motion of the atoms' [of my body] according to the laws of nature," i.e.,
who exercises my free will in the face of the laws of nature. Schrodinger's
answer to the question "What is this 'I'?" is that

If you analyze it closely you will, I think, fmd that it is just a little bit more than a
collection of single data (experiences and memories), namely that canvas upon which
they are collected. And you will, on close introspection, find that, what you really mean
by 'I' is that ground-stuff upon which they are collected. You may come to a distant
country, lose sight of all your friends, may all but forget them; you acquire new friends,
you share life with them as intensely as you ever did with your old ones. Less and less
important will become the fact that, while living your new life you still recollect the old
one. 'The youth that was I,' you may come to speak of him in the third person, indeed
the protagonist of the novel you are reading is probably nearer to your heart, certainly
intensely alive and better known to you. Yet there has been no intermediate break, no
death. And even if a skilled hypnotist succeeded in blotting out entirely all your earlier
238 S. S. SCHWEBER

reminiscences, you would not find that he had killed you. In no case is there a loss of
personal existence to deplore. Nor will there ever be.

Brandeis University

NOTES

1 F. Nietzsche, On the Genealogy of Morals, II. 13. See Basic Writings of Nietzsche,
translated and edited by W. Kaufmann (New York: Random House, Modern Library
Giant, 1968).
2 B. Barnes, Scientific Knowledge and Sociological Theory (London: Routledge and
Kegan Paul, 1974), p. 1.
3 G. Btlhme, W. van der Daele, and W. Krohn, 'Finalization in Science', Soc. Sci. In/. 15
(1977), pp. 307 -330.
4 See, in this connection. W. Stegmuller, The Structure and Dynamics of Theories (New
York: Springer Verlag, 1976).
5 R. Hersch, 'Reviving the Philosophy of Mathematics', Advances in Mathematics 31
(1979),pp.31-50.
6 E. Schrtldinger, What Is Life? (New York: Macmillan Co., 1945).
C. F. v. WEIZSACKER

THE UNITY OF NATURE

"In my beginning is my end." Thus begins T.S. Eliot's philosophical poem


East Coker, the second of his Four Quartels [1]. The application to our pre-
sent topic: philosophy and, I think, fundamental science move in circles or, if
the hope for progress is not delusive, in spirals. When we begin to speak about
language, we have been speaking, i.e., we have been using language, for years.
When we decide to learn scientifically from experience, we are already in pos-
session of the human experience of how to learn from individual and common
experience. Language and experience, science and philosophy are parts of hu-
man cultural history, the beginning of which is beyond our memory. And when,
on the other side, an end is achieved, say a scientific theory is completed,
one of its tests is its semantic consistency, that means its ability to justify
precisely that language and to explain precisely that experience which had
initially been needed to endow its concepts with an understandable meaning.
This lecture intends to describe a scientific program which cannot possibly
be carried out by a single person in a lifetime. The program is rather a theory
on what fundamental science has been doing since its Greek beginnings, and
how it ought to proceed now. The presentation in one lecture must be elliptic.
Ten years ago I described the program in a book which is now available in F.
J. Zucker's excellent English translation under the title The Unity ofNature
[2]. For details I must refer to the book or to later publications. My lecture,
too, will move in a circle, going through four sections: (1) philosophical
reflection, (2) a reconstruction of theoretical physics, (3) a program for
future work, (4) philosophical reflection.

1. "PHILOSOPHICAL REFLECTION

The term 'philosophical reflection' describes the fact that philosophical


thought refers to thought which existed before philosophy. Reflection,
literally understood, means mirroring. Our day's philosophy of science
reflects on the science which it fmds in existence; it reflects on its meaning,
on its rules, on the preconditions of its possibility. I shall try to reflect on the
philosophy of science which I find in existence.
Modern philosophy of science has, I think, its most important root in the
239
R. S. Cohen and M W. Wartofsky reds.), Physical Sciences and History of Physics, 239-254.
© 1984 by D. Reidel Publishing Company.
240 C. F. v. WEIZSACKER

conceptual problems created by the progress of theoretical physics. Ernst


Mach, a first-rate physicist and a pioneer in the history of physics, criticized
Newton's a priori presumptions and their modification in Kantianism, thus
paving the way for Relativity; Mach also criticized the a priori presumptions
of atomism and thus created the intellectual atmosphere in which the
Copenhagen version of quantum theory could grow. Mach's emphasis was on
experience as against apriorism, on phenomena as against ontological pre-
judice concerning reality. It is pathetic but it ought perhaps not to surprise us
that he failed to recognize the real progress of theoretical physics in his later
years: indeed, relativity and the revival of atomism. Anti-dogmatism can itself
become a blinding dogma.
In order to be brief I must present my personal view on the more recent
development of the philosophy of science without justifying it with a cohe-
rent chain of arguments. My view is mainly determined by the fact that I am
an active theoretical physicist. It cannot be denied that recent philosophy of
science has not contributed to actual progress in physics. I think that this
failure cannot be excused by the remark that philosophy is reflection rather
than a positive science. The problems on which this philosophy reflects are, I
think, simply not the problems which have turned out to be relevant in the
progress of physics. To give one example: the feud between logical positivism
and realism was applied to the interpretation of quantum theory after the
event; but no empirically relevant progress in or beyond quantum theory has
been produced by this debate. Of course I am aware of Bell's theorem and the
experiments connected with it. But these highly interesting results have
merely clarified and corroborated quantum theory. They have shown once
more that in this particular field, no progress in or beyond quantum theory
was needed. I suspect the problems of this debate are not the real problems of
physics. I shall return to this question in my second section. Here, I shall
propose the view that these problems are not the true problems of a philoso-
phy of experience either.
Reading, e.g., such excellent authors as Carnap and Popper, I cannot
escape the impression that in their debate both apply some uncriticized a
priori views; Carnap on the meaning of experience, Popper on the meaning
of reality. I found it to show great progress when T. S. Kuhn approached
the question of the meaning of empirical science empirically, which
means by a historical study. He discovered the phenomenon of recurrent
scientific revolutions, a structure which no logical positivist would have
predicted from his presuppositions on experience. Popper, it is true, has been
able to incorporate the phenomenon into his idea of successive, falsifiable
THE UNITY OF NATURE 241

hypotheses about reality. Yet I think that neither Kuhn's nor Popper's
descriptions stay close enough to real historical progress in basic natural
science, i.e., in theoretical physics. The salient point in this progress, I think,
was best described 15 years before Kuhn by Heisenberg [3] who could rely
on his own experience in creative theory shaping.
Heisenberg describes the progress of theoretical physics as a sequence of
closed theories ('abgeschlossene Theorien'). Classical mechanics, Maxwell's
electrodynamics, special relativity, and quantum mechanics are examples of
closed theories. A closed theory cannot be improved by small changes. But a
closed theory need by no means be the final truth. It cannot be reformed,
and for its field of applicability it needs no reform. But it can be replaced by
a new theory through a revolution. The new theory may use quite different
concepts, but it will contain in a sense what is not quite simply the old theory
as a limiting case and thereby explain its success and delimit its field of
applicability.
So far, this is only a (very abbreviated) description of historical facts. The
philosophical problem of physics can only now be formulated with sufficient
precision: how are closed theories possible? We will completely miss the
meaning of this question unless we open our minds to the overwhelming sense
of wonder if not awe justified by the fact that such theories exist and are
successful. For quantum theory, e.g., it is probably not an exaggeration to say
that its basic assumptions can be expressed in one page of print (for a mathe-
matically trained reader), but that there have been up till now something
like a billion empirical cases to which it can be applied successfully without
a single empirically confirmed counter-example. No methodological concept
(such as Mach's economy of thought) can explain this. The physicist can-
not escape speaking a realistic language. When tens of thousands of people
have travelled around the earth, we simply say: the earth is a ball. Accord-
ingly we simply say: nature obeys the quantum laws. Yet there is a difference
in apprehension. We can easily visualize the structure of a ball. How can we
imagine the meaning of quantum laws? Why should they hold? Here we seem
to have adopted realism without an idea of what reality means.
The problem is not new. Kant discovered it in the examples of Euclidean
geometry and Newtonian mechanics. Why should our experience strictly obey
Euclid's and Newton's laws, while we see no way of understanding the neces-
sity of these laws by any strict logical conclusion drawn from experience?
Kant's hypothetical answer is: these laws must be preconditions of experience,
i.e., conditions which must hold if experience is to be possible at all. This is
what is currently called subjectivism. Kant says: the conditions for the
242 C. F. v. WEIZSACKER

possibility of experience are at the same time the conditions for the possi-
bility of the objects of experience. To use the modern lin~uistic jargon:
objects of which we can speak are by necessity objects of which we can
speak; they belong to the field of possible experience. The philosophical pro-
blem here lies in the meaning of the word 'possible'. I shall return to this
later. The details of Kant's theory are now obsolete. But the program which I
propose is inspired by Kant's idea. The program says: we may try to recon-
struct the theories of physics from assumptions about the preconditions of
experience, assumptions which we try to choose as simply as possible. That
experience is possible, is presupposed as a fact; we ask under which con-
ditions it is possible.
The objection is natural: Kant has failed, why should we succeed? How
can we hope to deduce Newtonian mechanics, the two laws of thermo-
dynamics, Maxwell's equations or the existence of a Minkowski space from
preconditions of experience? But this objection is too simple. Evidently there
is no hope of deducing the many special theories such as mechanics, optics
etc. directly from philosophical principles. But physics has now reached a
high degree of unity. The laws of (relativistic) quantum theory seem to be
universal. They suffice for deducing classical physics and chemistry. Hope
even exists of deducing the system of elementary particles from fairly simple
symmetries. Thus the question is only whether these few universal laws can
be philosophically interpreted. In this sense the unity of nature would be the
clue to the validity of physics. The term 'unity of nature' here expresses the
fact that coherent experience is possible; it expresses this in the objective
language of the physicist.

2. A RECONSTRUCTION OF THEORETICAL PHYSICS

What I describe is a program, not an accomplished formal system. By 'recon-


structing' a theory, I mean, to a first approximation, finding a set of simple
and adequately plausible axioms for it. Two questions arise at the outset:
(a) Which theories do we want to reconstruct?
(b) What are adequately plausible axioms?
Ad (a): We may say that theoretical physiCS today consists of two inde-
pendent fundamental theories: quantum theory and relativity, and the as yet
incomplete third theory of elementary particles. As an additional, logically
independent theory we may add statistical thermodynamics. The mutual
relationship of these theories is somewhat complicated. Relativity is the
theory of the space-time continuum. Quantum theory can be presented as
THE UNITY OF NATURE 243

'abstract' quantum theory; in this form it does not refer to a space-time


continuum, but it needs an 'absolute' time coordinate. If we take over from
relativity no more than the existence of a three-dimensional Euclidean space
in which all objects move (that is the Galilean limiting case of special rela-
tivity), and if we accept the elementary particles as empirically given, we can
build up the non-relativistic quantum theory of matter, which is essentially
sufficient for explaining classical mechanics and, with thermodynamics, also
chemistry, hence - if physicalism should turn out to be correct - even all of
biology. It is the reconstruction of statistical thermodynamics and of abstract
quantum theory for this purpose, at which the present section aims. The
following section will refer to relativity.
Ad (b): A mathematical formalism (say the geometry of Minkowski space
or of a Hilbert space) is not yet a physical theory. It needs a physical seman-
tics: an interpretation of the mathematical concepts in terms of experience,
as described in everyday language. Axioms oUght to be plausible, if expressed
in this language. The high-strung ambition of a reconstruction as indicated at
the end of the preceding section would be to confine the axioms to state-
ments which express no more than the preconditions of all experience.
A precondition of experience is time. Experience may be defined as learn-
ing from the past for the future. In this statement, the terms 'past' and
'future', and hence, by unstated implication, 'present', are used as though
they were understood. They are not explained, but taken over from everyday
language, and they are used for explaining such a fundamental term as
'experience'. When I speak of 'time' or 'time-structure' without further quali-
fication, I always mean this triad of the so-<:alled modes of time: present,
past, future. Whether time, in addition to this, is correctly measured by a real
coordinate, is in my approach a point of secondary importance. It is essential
for my approach to see that no empirical concept has any meaning unless we
implicitly presuppose our understanding of the modes of time, and that this
direction of dependence cannot be inverted. There is no way of eliminating
present, past and future from physics or of explaining them by terms which
do not implicitly presuppose them; if we try that, we end up eliminating all
meaning from physics.
This is a strong statement of which, in my experience, a born physiCist can
only become convinced by unsuccessfully trying for years to evade it. I may
perhaps indicate how it became clear to me. In 1939 I wrote a paper on the
second law of thermodynamics [4]. Let me first quote a remark by Einstein
[5]: "A theory is the more impressive the greater the simplicity of its pre-
mises is, the more different the kinds of things it relates, and the more
244 c. F. v. WEIZSACKER
extended is its area of applicability. Hence the deep impression which classi-
cal thermodynamics made upon me. It is the only physical theory of universal
content concerning which I am convinced that, within the framework of the
applicability of its basic concepts, it will never be overthrown ( for the special
attention of those who are skeptics on principle)." Einstein wrote this in
1948, thus I could not know it in 1939. But I had a strong feeling in the same
direction. My problem was how a reversible basic theory can yield irrever-
sible processes when statistics, Le., the concept of probability, is applied to it.
Reading the classical papers by Gibbs and the Ehrenfests, I came to the con-
clusion that 'probability', in the sense in which it is used in statistical thermo-
dynamics, can only apply to the future, not to the past. The past is .factual, it
can be known, the future consists in possibilities; only to possibilities can we
ascribe the concept of probability in an objective sense. If this is correct, the
most universal law of physics rests on the modes of time.
This would be a topic for a complete lecture. Together with M. Drieschner,
I am preparing a book on the concept of probability [6] , its connection with
temporal logic, with thermodynamics, evolution, and quantum theory. I
shall just say here that we define probability as the expectation value of a
relative frequency; and that by 'expectation' we mean expectation, that is we
mean the cognitive anticipation of future events. This is the concept of pro-
bability which we also use in the reconstruction of quantum theory.
Abstract quantum theory maintains that the possible states of any physical
object can be represented by rays in a Hilbert space. The connection of this
mathematical structure with experience is given by the concept of probability.
The values of probabilities are defined by the Hilbert metric: in a given state
of the object, the probability of finding another possible state of that object
by an appropriate experiment equals the absolute square of the inner product
of two normalized vectors belonging to the two states. The probability thus
defined is non-classical in the following sense. According to classical, say
Kolmogorov's, axioms, probability is a real non-negative valuation of a set
of possible events; this set has the lattice structure of a Boolean algebra
because the events can be expressed by statements which obey the classical
logic of propositions. J. von Neumann first pointed out that the set of
quantum events is the non-Boolean lattice of the linear subspaces of
Hilbert space. Von Neumann spoke of it as a non-classical quantum logic.
There has been some disagreement on whether this lattice deserves the
name of a logic. In my view the solution is that quantum logic is a special
version of a temporal logic. To ascribe a probability p to a possible event
means making a statement about the future: if you repeat the relevant
THE UNITY OF NATURE 245

experiment many times, you can expect a relative frequency p for the event
it question.
The problem of a reconstruction of abstract quantum theory is to find
plausible axioms from which this structure of the lattice of events with its
probabilistic interpretation can be deduced. This is a well-known field of
study today, called quantum axiomatics. The main progress in this field has
been made by J. Jauch [7] and his school. M. Drieschner [8] took a further
step by explicitly using the ideas of temporal logic. I have made several
different attempts at such a reconstruction, some of which have not yet been
published [9]. Here, I shall give an outline of one of them.
The reconstruction is done in several steps which can be characterized by
the following concepts:
(a) temporal logic of alternatives,
(b) indeterminism,
(c) objects,
(d) symmetry and real quantum theory,
(e) dynamics and complex quantum theory.
I can only indicate the outline of the argument.

(a) Alternatives. We describe all possible observations as decisions of n-fold


alternatives. n here means either an integer> I or denumerable infinity. An
n-fold alternative means a set of n formally possible events which fulfIl the
following conditions:
(1) The alternative can be decided; i.e., a situation can be created in which
one of the possible events becomes an actual event and henceforth a fact
(transition from future possibility through present actuality to past facticity).
We then say that this event has taken place.
(2) If one event in the alternative has taken place, none of its other events
have taken place. The events of an alternative are mutually exclusive.
(3) If the alternative has been decided and all of its events except one have
not taken place, this last one has taken place. The alternative is defined as
complete.

(b) Indeterminism. We postulate: For every alternative there are well-


defined situations in which none of its elements is a fact. 'There are' means:
such situations are possible in principle or, as I say, are formally possible. A
complete theory of an alternative or of an object will describe all that is
formally possible with respect to this alternative or object. It will hence
have to include well-defined situations in which the alternative is undecided.
246 C. F. v. WEIZSACKER

We further postulate: For any well-defined situation belonging to an


alternative there exists a probability function over the possible events of the
alternative, expressing the probability of finding the relevant event, given
the situation in question. This seems to be the minimum assumption under
which we would consider the situation as well defmed.
The postulate of indeterminism expresses the non-classical nature of quan-
tum probabilities and may be considered as the basic assumption of quantum
theory. We have not formulated it as some kind of a metaphysical truth. But
if it holds true, the assumption that an alternative is always objectively
decided and that probability only expresses incomplete knowledge, cannot be
upheld. This will be settled once we accept the postulate of symmetry. (See
(d).)

(c) Objects. I give a very abstract definition of an object which I can quote
verbally but not fully explain with the necessary brevity. The set of all well-
defined situations belonging to a given alternative A is called the set of the
formally possible properties of the object which belongs to A. I shall not
quote the precise definition of the term 'belonging to'. In the final quantum
theory, the alternative A will mean a maximal observable of its object. It is
seen that I do not start out with a so-called realistic assumption of objects
with properties, but by assuming laws for alternatives decidable in well-
defined situations; laws which finally justify the use of the term 'object'
precisely for that degree of approximation in which the situation can be
described as well defined.
We postulate the existence of a probability function for every point of
formally possible states of an object: P (x, y) is the conditional probability of
findingy in an appropriate experiment when x is the case.

(d) Symmetry and Real Quantum Theory. We postulate a law of symmetry


for the probability function: Any state of an object is an element of an alter-
native which is equivalent to the defining alternative A of the object. Two
alternatives A and B are called equivalent if there is a mapping of the set of
states of the object onto itself which preserves the probability function and
transforms A into B. Hence an object can be equally defined by anyone
of its alternatives.
This postulate defines the approximation within which we can speak of
separate objects. Where it does not hold, the given alternative A is to be con-
sidered as a part of a larger alternative belonging to a compound object of
which the object defined by A is an interacting part. In fact all known objects
THE UNITY OF NATURE 247

interact with some other objects, since without interaction we could not know
of them. Hence the concept of an object is necessarily no more than an
approximation.
The symmetry of an object can be expressed by a group of transformations
acting on its space of states. We can represent this group in a real vector space.
The probability function serves to define a real positive metric in this space
which is kept invariant by the group. Hence the group must be orthogonal.
Skipping over a few arguments I assume it to be a simple orthogonal Lie group.

(e) Dynamics and the Complex Vector Space. We postulate: The evolu-
tion of all states in time must be described by a one-parameter subgroup of
the symmetry group, the parameter being the time t.
Here we presuppose that time can be represented by a real continuous
parameter. This is perhaps not the final truth, but it is the assumption on
which the existing quantum theory rests. For the further assumption that
the dynamical evolution preserves the probability function we can advance a
'Darwinian' argument: only those states which retain their identity under
motion will be observable; and this identity is defined by their statistical
relations to other states.
Being orthogonal, the generator of the dynamical one-parameter group,
the 'Hamiltonian', H, can be near-diagonalized into a matrix consisting of 2 X
2-matrices of the form

along the diagonal, and zero elsewhere. Hence the group can equally well be
represented in a complex vector space with half the number of dimensions
and the dynamical operator H a diagonal matrix with elements e - ;W;t. In
this space the symmetry group turns out to be both orthogonal and symplec-
tic, hence unitary.
Thus we reconstruct abstract quantum theory. In my view, the two main
assumptions are the postulates of indeterminacy and of symmetry. Indeter-
minacy, I feel, reveals the meaning of an open future, hence a fundamental
structure of time. The symmetry of an object means that the alternatives can
be introduced step by step, distinguishing the inner structure of an object
from its interaction with other objects. If this process of successive approxi-
mation were impossible, we would have to decide all alternatives at once,
which means that human, i.e., finite knowledge would be impossible.
248 C. F. v. WEIZSACKER

3. PROGRAM FOR FUTURE WORK

Concrete quantum theory, the specific theory of the real objects of nature,
presupposes the existence of a three-dimensional'position space and of special
objects, called particles, in it. Position space belongs, in the modern view, to
relativity. Hence it seems natural, as a next step, that we try to reconstruct
at least special relativity.
Here the essential idea is to deduce relativity from quantum theory by a
simple assumption of the symmetry of the dynamical law. My first papers on
this idea were published in 1955 [10] and 1958 [11], the latter with E.
Scheibe and G. Stissmann. My book The Unity of Nature contains a brief
account (Section II. 5). Since 1968, D. Finkelstein [12], quite independently,
has pursued a similar line. In the seventies I continued this work with several
collaborators, in the first place with L. Castell. The four-volume Quantum
Theory and the Structure of Time and Space (1975-1981) [13] contains
some of our results. Let me be vain enough to include a pleasant personal
recollection here. When I first met David Finkelstein, in 1971, neither of
us knowing much of the work of the other, I told him that I thought
Minkowski-space might be deduced from the quantum theory of a binary
alternative. He said: "You are the only man in the world to say such a thing.
Of course you are right." It is clear that he said so because I was not the only
man of that description; he was the other one.
The first idea was what I would call radical atomism. Traditional atomism
contains the slight inconsistency that it gives no answer to the question why
atoms should at all be indivisible. Chemical atoms are extended; why should
there not be parts to them? Present-day elementary particles mostly have
finite mass; why should this mass not be divisible? Can the sequence atom-
lepton and baryon-quark come to a logically necessary end? In quantum
theory, the composition of objects is described by forming the tensor product
of the Hilbert spaces of the composing parts. The smallest possible factor in
such a product is the two-dimensional vector space. (A one-dimensional
factor only adds a universal phase which carries no information.) This two-
dimensional space corresponds to the binary alternative, to a simple yes-no
decision. I assume that there exists a basic physical object which admits only
binary alternatives as observables; I call it the ur (German: Ur-Objekt, Ur-
Alternative). This is a falsifiable hypothesis in Popper's sense only if it is
understood to defme the symmetry of the dynamical law; a Hilbert space
taken alone, without time-dependence, can always be factored into binary
spaces in an infinite number of ways. Yet it is a common assumption in parti-
THE UNITY OF NATURE 249

cle physics that in a composite object, the law of symmetry for the free
motion of the parts also applies to their interaction; e.g., we assume Poincare-
invariance for the equation governing all free motions of particles and we
assume it for the equation of their interaction, too. This same logical struc-
ture we now apply to the theory of the ur. We fmd the law of symmetry for
the free motion of urs, and then we demand the same symmetry for their
interaction. Hereby we shall find a fairly simple hypothetical symmetry group
for all of physics. The hope is that this group will contain both the Poincare
group which defines relativity, and the internal symmetry characteristic for
elementary particles.
This investigation goes through several steps, and it is as yet unfmished. In
indicating the results already achieved or, as we hope, near at hand, I must for
just a few minutes use more technical language than has been used in the rest
of this lecture.
We must distinguish between metrical and dynamical symmetry. Metrical
symmetry is the symmetry of quantum theory which keeps the Hilbert-space
metric invariant. Its group consists of unitary and anti-unitary transformations
and the discrete element of complex conjugation. For the single ur,. the unitary
transformations -leaving the phase transformation aside - make up the group
SU (2). SU (2) is locally isomorphic to SU (3). Ifwe now apply the principle
that every object composed of urs must admit the same symmetry group as the
free ur, we conclude that all objects of physics must admit SO (3), that is the
group of rotations in a locally Euclidean three-dimensional real space. Thus the
universal three-dimensional position space seems to be a natural and necessary
consequence of the ur hypothesis. This remark was my starting point in 1955.
The second step consists in considering complex conjugation. In the two-
dimensional space, this is a non-linear transformation. It is mathematically
convenient to express it linearly in a doubled space, that is, in four complex
dimensions. The physicist would say that we thereby introduce an anti-ur,
similar to an anti-particle. This step was already considered in our 1958
paper, but the form we now use was given by Castell in 1975 [14].
The third and most important step is the search for the full dynamical sym-
metry, that is for the group which keeps the equation of motion invariant.
Castell showed that in his representation this group is SU (2, 2) which is
locally isomorphic with SO (4, 2), the conformal group of special relativity.
This is the shape in which we present the idea today on which I agreed with
Finkelstein long ago: that special relativity is a consequence of the quantum
theory of the binary alternative.
The fourth step is rather a long march. I call it the realization of the ur
250 C. F. v. WEIZSACKER

theory. We must try to frame explicit models showing how particles can be
built up from urs. Minkowski- or de Sitter-space is introduced as a homo-
geneous space of the relativistic group. The single ur contains no more
information than one bit. Hence it cannot be localized in this space. A super-
position of many urs, however, can be more nearly localized. The state-space
of one particle is, according to Wigner, an irreducible representation of the
Poincare group. Such representations can be constructed in the tensor space
which consists of all tensors of any finite rank over the four-dimensional
vector space in which we represent the single ur and anti-ur.
Here we must first resolve an apparent paradox. Abstract quantum theory
uses an absolute time coordinate. How can it become relativistic, even by
using a specialization like the ur hypothesis? The mathematical answer is: The
non-compact elements of the relativistic group, as represented in the tensor
space, do not keep the number of urs constant, while, of course, they always
transform a binary alternative into another binary alternative. Every ur is a
binary alternative but not every binary alternative is an ur. A Lorentz trans-
formation transforms a single ur into a linear combination of many urs. This
means that the concept of ur is relative to the chosen frame of reference. In
terms of physics, this means: The time-coordinate of abstract quantum theory
is the proper time ofasingle observer. A relativistic transformation connects two
different observers. Different observers will use different definitions of the ur.
Castell has shown that the subspace of symmetric tensors represents just
one massless particle of every helicity. This corresponds to ascribing Bose
statistics to the ur. In order to represent many-particle systems and massive
particles we must use tensors of all symmetry classes. Castell and his colla-
borators (Jacob [15], Heidenreich [16], Kuenemund [17]) have studied
Para-Bose statistics. Quite recently, I have become convinced that Para-Bose
operators can be represented in the tensor space of the ur, and that this
representation even exhausts the relevant information contained in this space.
We are now studying these representations.
I have only been able to give a brief introduction to work actually in pro-
gress. I have had to omit the incomplete attempts which have been made to
approach the theory of real elementary particles. I feel that nevertheless I was
obliged to say at least this much, in order to show that the hope of fully
uniting the basic theories of physics is not just a philosophical chimera but
may well end up by becoming solid theoretical physics.

4. PHILOSOPHICAL REFLECTION
The philosophy implied in the preceding considerations is, I am afraid, some-
THE UNITY OF NATURE 251

what different from all existing philosophies of science. It is inspired by the


realism of physicists in speaking of the unity of nature, and not primarily of
the scientific method; yet it is not the usual kind of realism since it does not
presuppose definite characteristics of reality, and it argues from the condi-
tions of knowledge in the way we have all learned from Niels Bohr. It is
inspired by empiricism and by Kant alike, considering experience as the
central concept and putting all its results to empirical test; but it is not
empiricism, since it assumes that the validity of natural law is not a logical
consequence of special experience but a precondition of all experience; and
it is not Kantianism since it realizes this assumption not by a priori state-
ments but hypothetically. It is inspired by Platonism in relying on universal
mathematical structures, but it is not Platonism since it gives time the first
priority. It is inspired by Kuhn in learning from history what empirical
science means, but it follows Heisenberg rather than Kuhn in considering the
possibility of closed theories as the central problem of a philosophy of
science.
In the beginning, I said that the controversy between positivism and
realism misses the real problems of both physics and philosophy. About my
attempted reconstruction of physics, it is difficult to say whether it is
philosophy or physics. I like this difficulty. The separation of empirical
science from philosophy is good for the Kuhnian periods of normal science; if
you possess a paradigm, philosophy questioning it is just a nuisance. In
revolutions, philosophical thought is absolutely necessary for science.
I mentioned Niels Bohr, and I hope that my reconstruction is in the
Copenhagen spirit. Bohr's position has erroneously been subsumed under the
term 'positivism'. He studies the conditions for knowledge. Knowledge means
that somebody knows something. The emphasis on the role of the observer is
not 'subjectivism', whatever this word may mean. We all know that two
observers seeing the same mountain from diffeIent sides see different phe-
nomena; but we can hope to predict what will be seen by different observers.
The lack of so-called objective reality in quantum theory is, I think, explained
by the consideration of time. The past is factual. The mistake of average
realism is to think that future events must be equally factual; I wish to
emphasize that Popper [18], taking the reality of time as the cornerstone
of realism, avoids this mistake. The future is possible. It can be shown that
different observers can correctly ascribe different probabilities to the same
possible event since they each meaningfully consider it as an element in a
different statistical ensemble. All apparent paradoxes of the reduction of
wave functions, of the Einstein-Podolsky-Rosen experiment etc., disappear
252 C. F. v. WEIZSACKER

completely once we realize that the wave function expresses an evaluation of


future possibilities, given some factual knowledge of the past.
An interesting so-called paradox is known by the name of 'Wigner's friend'
[19]. It considers the state of mind of an observer and intends to show that
quantum theory cannot apply to consciousness. I think the argument is
erroneous. Abstract quantum theory, as I have presented it, does not refer to
anything other than decidable alternatives. If the state of mind of an
observer, whether he is different from myself or whether I am the observer,
,can be tested unambiguously, then it can also be predicted with probability
and is in principle subject to quantum theory. Quantum theory would be
fully consistent with a monism which would say: soul or mind or feeling,
culminating for the time being in consciousness, is the essential quality of na-
ture; to the extent that it can be subjected to decidable alternatives, it obeys
quantum theory and hence will also show all the characteristics of matter.
Let me explain this idea by drawing a comparison with Cartesian dualism
and its - I think insurmountable - difficulties. When Descartes had finished
his Meditationes de prima philosophia, he sent them to several eminent
philosophers of his time, among them, Thomas Hobbes. Hobbes's answer
amounted roughly to his saying: "In his definition of res extensa and res
cogitans, M. Descartes is begging the question. He defines them such that each
of them is deprived of the quality defining the other. I cannot see why an
extended substance should not be able to think." This, I think, is the valid
objection of a materialist to Descartes' method. But while I agree that there is
no evident reason why real matter primarily defined by us as extended,
should not in fact have the ability to think, I cannot see an evident reason
why it ought to have this ability either. Materialistic monism in this sense
remains mere guesswork, founded no better methodically than Cartesian
dualism. Here the subjective predilection of one philosopher stands against
the subjective predilection of another philosopher.
But if you tum the argument around and use my description of quantum
theory, you may come to a more stringent conclusion. Let us, hypothetically,
begin by ascribing the qUality of cogitatio to the basic 'substance' of the
world (knowing, of course, ,that the concept of substance needs further analy-
sis as well). Cogitatio is Descartes' term for what I tentatively circumscribed
by the words 'soul' or mind' or 'feeling' or 'consciousness'; 'subjectivity' is
another philosophical term for it. Now insofar as cogitatio can be subjected
to an analysis in terms of decidable alternatives, it will obey the laws of
quantum theory, hence it will - if my theory of un is correct - necessarily be
Lorentz-invariant, i.e., it will admit a spatial description: it will by necessity
THE UNITY OF NATURE 253

appear as extended. Matter, using this description, would be mind as


perceived by a mind. To fall back on linguistic jokes: thought can think
extension, it can even think of itself as extended; extension, however, cannot
be extended to become thought. Or again, in other language: thought is re-
flective, it can refer to itself as being different; extension is not reflective, it
would need the quality of thought as an addition, not as a consequence of
its own nature. It is clear that such a theory would need possibility or poten-
tiality as a necessary component. If the 'substance' of this description
is not necessarily thinking, it would be potentially thinking, and the reali-
zation of this potentiality would be evolution.
I offer these last ideas as a philosophical hypothesis, not as a stated truth.
It is not a necessary consequence of quantum theory, but it would be consis-
tent with quantum theory.
I have finished my presentation. Evidently, it was not yet a philosophy,
but a philosophical inquiry into special problems. But I think Of these further
questions: What can we say of reality when we do not confine ourselves to
decidable alternatives? How is empirical theory related to moral responsibility,
to social relations, to artistic intuition, to religious experience? How is
rational consciousness related to love or to meditative awareness? What is the
universe? Where do we stand in evolution? Is time an ultimate?
T. S. Eliot's poem ends with the line: "In my end is my beginning."

Max Planck-Institut, Stamberg

REFERENCES

[1) Eliot, T. S.1944. Four Quartets (London: Faber and Faber).


(2) Weizsiicker, C. F. v. 1980. The Unity of Nature, transla. by F. J. Zucker (New
York: Farrar, Straus, Giroux); German Edition: Die Einheit der Natur (Munich:
Hanser, 1971).
(3) Heisenberg, W. 1948. 'Der Begriff "Abgeschlossene Theorie" in der modemen
Naturwissenschaft.' Dialectica 2 331.
(4) Weizsiicker, C. F. v. 1939. 'Der zweite Hauptsatz und der Unterschied von Ver-
gangenheit und Zukunft.' Ann. d. Physik 36 275; reprinted in The Unity of Nature
(see (2).
(5) Einstein, A. 1949. 'Autobiographical Notes.' In P. A. Schi1pp (ed.), Albert Einstein
Philosopher-Physicist p. 33 (Evanston, m.: Library of Living Philosophers).
(6) Weizsiicker, C. F. v. 1973. 'Probability and Quantum Mechanics.' Brit. J. Phil. Sci.
24321.
(7) Jauch, J. M. 1968. Foundations of Quantum Mechanics (New York: Addison-
Wesley).
254 C. F. v. WEIZSXCKER

[8a] Drieschner, M. 1970. 'Quantum Mechanics as a General Theory of Objective


Prediction.' Dissertation, University of Hamburg.
[8b] Drieschner, M. 1979. 'Voraussage-Wahrscheinlichkeit-Objekt tiber die begrifflichen
Grundiagen der Quantenmechanik.· In Lecture Notes in Physics (Berlin: Springer
Verlag).
[9] Weizsacker, C. F. v. 1979. 'A Reconstruction of Quantum Theory.' See [13],
Volume III. Unpublished: Abstrakte Quantentheorie (1974); Temporal Logic and
a Reconstruction of Quantum Theory, preprint (1978); A New Reconstruction
of Quantum Theory, preprint (1980).
[10] Weizsacker, C. F. v. 1955. 'Kompiementaritiit und Logik.' Die Naturwissenschaften
42 521 and 545. Reprinted in C. F. v. Weizsiicker, Zum Weltbl7d der Physik
(Stuttgart: Hirzel, 1957).
[11] Weizsiicker, C. F. v., E. Scheibe and G. Siissmann. 1958. 'Kompiementaritiit und
Logik III. Mehrfache Quanteiung.· Z. f. Naturforschung 13a 70S.
[12] Finkelstein, D. 1968. 'Space-Time-Code.' Phys. Rev. 184 1261.
[13] Castell, L., M. Drieschner, and C. F. v. Weizsacker (eds.). 1975-1981. Quantum
Theory and the Structures of Time and Space. (Munich: Hanser. Papers from the
Conference on Quantum Theory and the Structures of Time and Space. Vol. I
(1975) = 1st Conf., Feldafmg, 1974; Vol. 11(1977) = 2nd Conf., Feldafmg, 1976.
Vol. III (1979) = 3rd Conf., Tutzing, 1978. Vol. IV (1981) =4th Coni'.)
[l4] Castell, L.1975. 'Quantum Theory of Simple Alternatives.' In [13], I.
[IS] Jacob, P. 1978. 'Konform invariante Theorie exklusiver Elementarteilchen -
Streuungen bei grossen Winkeln.' Dissertation, Starnberg.
[16] Heidenreich, W. 1901. 'Die dynamischen Gruppen, SOo (3.2) und SOo (4.2) als
Raum-Zeit-Gruppen von Elementarteilchen.' Dissertation, Starnberg. [20].)
[17] Kuenemund, T. 1981. Diplomarbeit, Starnberg.
[18] Popper, K. 1974, 1976. Unended Quest. An Intellectual Autobiography, Chapter
28 (on talks with Einstein) (London, Glasgow, La Salle, lli.: Open Court).
[19] Wigner, E. P. 1961. 'Remarks on the Mind-Body Question.' In The Scientist
Speculates, ed. by I. J. Good (London: Heinemann). Also in E. P. Wigner, Sym-
metries and Reflections. Scientific EaSllys (Cambridge, Mass.: MIT Press, 1967).
INDEX OF NAMES

Abele, J. 53,55 BrogUe,L.de 15,20,21,22,43,44,45,


Alembert, J. L. R. d' 206 53
Anderson, C. D. 12,40 Brown, Robert 6
Arago, D. J. F. 35 Bruno,G. 5
Araki, H. A. 146, 158 Biichel, W. 186,187,188
Archhnedes 116,130 Bunge,M. 28
Aristuchus 119 Burian, R. 232
Aristotle 4,5,6,41,44,47,116,117, Burtt, E. A. 115,117,120,140,141
120,123 Butts, R. E. 139
Arzelies, H. 187,188
Autrecourt, N. d' 5 Capek, M. 27,28,29-31,33,34
Avishai, Y. 57,74 Cunap, R. 147, 240
Cunot, S. 39,40,42,43,46
Bailey, C. 4,26,27 Cur, B. 113
Balzer, W. 188, 189 Cuter, B. 107,113
Bu-Hillel, Y. 55 Cassirer, E. 120, 141
Barnes, B. 238 Castell, L. 248, 249, 250, 254
Bauer, E. 28,38,45,51,52,55 Cauchy, A.-L. 204,206
Baumrin, B. 113 Cicero, M. T. 19
Bayes, T. 39,40 Clauser, J. 56
Bell, J. S. 48,56,240 Clausius, R. J. E. 206
Benioff, P. 73 Cohen, I. B. 142
Bergson, H. 19, 22, 24, 28, 40, 52, Cohen, R. S. 1,139,140
55 Colodny, R. G. 55
Blackwell, R. J. 229,230 Copernicus, N. 36,117,119,211
Bohm,D.15,18,20,28,48,56 Costa de Beauregud, O. ix, 55, 56
Bohme, G. 186, 188, 197, 234, 235, Cox, R. T. 53,56
. 238 Crew,H. 140
Bohr, N. 10,12-13,17,27,211,221, Crick, F. 235
251
Boltzmann, L. 36,40,42,43,47,53, Dalton, J. 4,8
208 Danieri, A. 158
Bondi, H. 111,112 Davidson, D. 76-86, 88-90, 92
Born, M. 15,19,64, 150, 151 Democritus 3,4,5,6,25
Boutroux, E. 19 de Salvio, A. 140
Boyle, R. 5,234 de Santillana, G. 140
Brainud, A. J. 55 Descutes, R. 125,208,209,224,252
Brans, C. H. 100 Deser, S. 74
Bridgman, P. W. 220 Dicke, R. H. 100,107,113
Brillouin, L. 41,43,56 Dijksterhuis, E. J. 5
255
256 INDEX OF NAMES

Dingler, H. 161-162, 164, 186, 188, Fresnel, A. J. 35, 36


192 Frisch,O.von 13,23,27,28
Dirac, P. A. M. 16,19,28,37,40,45,
54,56 Gabor,D. 41
Drake, S. 115,124,125,127,135,139, Galileo 115-120, 122-128, 130, 131,
140,141,142 133, 134-137, 140, 141, 142,199,
Drieschner, M. 244, 245, 253, 254 200,208,209,211,220
Drude,P. 27 Gal.Qr, B. 55
Duhem, P. 3, 87 Gassendi, P. 4,5,19
Diisberg, K. J. 186 Geroch, R. 73
Gibbs,J. W. 36,39,55,209,213,244
Eddington, A. 15, 27 Gilbert, N. 120, 141
Ehlers, J. 57,74 Gilbert, W. 120,131, 133, 142
Ehrenfest, P. 244 GOdel, K. 112
Ehrenfest, T. A. see Ehrenfest-Affanas- Gold, T. 113
jewa, T. A. Good, I. J. 254
Ehrenfest-Affanasjewa, T. A. 244 Goodman, N. 75, 92
Einstein, A. 8-9, 11, 15, 18, 20, 27, 33, Green, G. 50, 54
35, 36, 38, 40, 42, 43, 44, 48, 53, Griinbaum, A. 57,74, 185, 187, 188
56, 57, 58, 59, 64, 69, 72, 73, 74, Gudehus, D. H. 113
111, 165, 208, 209, 211, 212, 221, Gutenberg, J. 211
243-244, 253, 254
Einstein-Podolsky-Rosen (EPR) para- Haag, R. 73
dox, etc. 38,45,48-51,251 Hadamard, J. 35, 55
Ekstein, H. 74 Hamilton, W. R. 71, 206
Eliot, T. S. 239, 253 Hanson, N. R. 113
Ellis, G. F. R. ix, 112, 113 Harrison, E. R. 108,112, 113
Elsasser, W. M. 44, 56 Havas, P. 73, 74
Emiques, F. 4, 26, 27 Hawking, S. W. 112
Epicurus 19, 20 Heckmann, O. 112,113
Espagnat, B. d' 48,49,54,55,56 Heelan, P. 34
Euclid 124,129,139,141,241 Hegel, G. W. F. 1,22
Euler, L. 204, 209 Heidenreich, W. 250,254
Evans, R. D. 27 Heintzmann, H. 187, 188
Everett, H., III 53,147,158 Heisenberg, W. 13, 15, 17, 26, 27, 33,
37,235,241,251, 253
Fantappie, L. 40, 55 Helmholtz, H. von 24
Farrington, B. 4, 27 Hempel, C. G. 78,88
Fermat, P. de 35 Heraclitus 207
Feyerabend, P. K. 141 Hersch, R. 238
Feynman, R. P. 45,49 Hesse, M. 137,143
Finkelstein, D. 248, 249, 254 Hintikka, J. 116,120-121
Fock, V. 44, 56 Hippocrates 52
Foster, L. 92 Hobbes, T. 252
Frank, P. 15,27, 176, 187, 188, 220 Holton, G. 200, 211, 216, 230
Franklin, B. 234 Hooker, C. A. 38, 55
Freedman, S. J. 56 Hoyle, F. 112
INDEX OF NAMES 257

Hume, D. 25, 90 Laue, M. von 15,27


Husserl, E. 30, 33, 140 Lavoisier, A. L. 25
Huygens,C. 9,14,27 Lebedev, P. N. 11
Lee, T.D. 53
Infeld, L. 64 Leibniz, G. W. 3
Leucippus 3
Jacob, P. 250,254 Levy-Leblond,J.-M. 187,188
James, W. 231 Lie, S. 63,74
Janich, P. 159, 164-168, 170, 172, Locke,J. 25
174-175, 180, 181, 184, 186, 187, Loinger, A. 158
188,192,196,197 London,F.38,45,51,52,55
Jaspers, K. 199 Longair, M. 112,113
Jauch, 1. M. 38,55,245,253 Lorentz, H. A. 4,35,36
Jaynes, E. T. 44,56 Lorenz, K. 189
Jeans, J. H. 20 Lorenzen, P. 162-163, 164,186, 188,
Jordan, P. 147 192
Joule, J. P. 67 Loschmidt, J. J. 39,52
Lucretius 5,6,19,20,27
Kambartel, F. 186,188 Ludwig, G. 50,55,147
KamIah,A.186,188,189
Kant, I. 24, 25, 68, 160-161, 163- McCrea, W. H. 106,113
164, 185, 186, 188, 192,241-242, Mach, E. 3, 24, 115, 142, 211; 240,
251 241
Kaufmann, W. 238 Mackey, G. W. 74
Kelvin see Thomson, William MacLachlan, I. 127,142
Kepler, J. 36, 227 McMullin, E. 139, 140
Klein, F. 63 McTighe, T. 120,141
Kolmogorov, A. N. 244 Malvaux, P. 53,55
Korner, S. 28 March,A. 28
Koyre, A. 115, 140, 141 Margenau, H. 17, 28
Krips, H. P. 38,55 Maxwell, J. C. 36,40,41,64,160,209,
Kristian, J. 112 241,242
Krohn, W. 238 Mehra, I. 74
Kronecker, L. 54 Melvin, M. A. 73
Kuenemund, T. 250,254 Mendeleev, D. I. 237
Kuhn,T.201,219,234,240-241,251 Metzner, A. W. K. 113
Meyerson, E. 4, 14, 19, 26, 27
Lakatos, I. 137,143,219,234 Michelson, A. A. 9,21,23,35
Lande, A. 14-16, 18, 19-20, 23, 26, Mill, J. S. 17
27,28,38,45,54,55,56,154 Minkowski, H. 36,49,53
Landolt, H. 11 Mittelstaedt, P. 186,187,188,189
Landsberg, P. T. 56 Mittelstrass, J. 186, 188
Laplace, P.-S. de 25,36,206 Moldauer, P. A. 37,50,55
Larmor, J. 35 M~ller, C. 74, 187, 188
Lasswitz, K. 4, 26, 27 Montagu, A. 141
Laudan, L. 185, 201, 208, 228, 229, Morrison, P. 113
230 Musgrave, A. 143
258 INDEX OF NAMES

Neumann,J. von 37,38,44,45,46,51, Reynolds, o. 9, 27


52,55,145,146,148,244 Rhine, J. B. 47
Newton, I. 3,5, 25, 30, 36,43, 71, 136, Riesz, M. 45,56
137, 143, 204, 206, 209, 210, 240, Ritz, W. 44, 56
241 Rosen,N. 56
Nietzsche, F. 232, 238 Rothe, H. 176, 187, 188
North, J. D. 111, 112 Russell, B. 10

Occam, W. see Ockham, W. Sachs, R. K. 112


Ockham, W. 94 'Sagredo' 118,122,123,124,125-127,
Onsager, L. 207 128, 131, 132, 141
Ostwald, W. 3 'Salviati' 117,118-119,120,122,123,
124, 125-127, 129-130, 131-132,
Palter, R. 28 133,134,135, 141
Pappus of Alexandria 121 Sandage 111
Parrnenides 4 Schaffner, K. 27
Peirce, C. S. 19 Schatz mann, E. 112
Penrose, R. 112 Scheibe, E. 248, 254
Perrin, J. 8, 27 Schelling, F. W. J. von 1,22
Pfarr, J. 186,189,191-197 Schild, A. 58, 74
Piaget, J. 25,28,29,30,31,32-33 Schilpp, P. 74, 253
Pitt, J: C. 139 Schlick, M. 2
Planck, M. 21,22,43,47,53,209,212 Schmidt, H. 47,56
Plato 116, 139, 140, 141 Schrodinger, E. 17, 18, 20, 21, 28, 35,
Podolsky, B. 56 37, 38, 46,48,50,51,53,56, 145,
Poincare, H. 10, 22, 24, 35, 36,53,55, 146,151,153,237-238
57,59,63,64,67,69,74, 206 Schiicking, E. 112, 113
Poianyi, M. 213, 225, 230 Schweber, S. S. 45,56
Popper, K. R. 14-16, 17, 18, 19-20, Schwinger, J. S. 45
23, 24, 26, 27, 28, 132, 202, 205, Sciama, D. W. 112
217,230,240-241,248,251,254 Segal, I. 73
Price, G. R. 47, 56 Sellars, W. 83,92
Prosperi, G. M. 50,56, 158 Settle, T. 115, 140
Shapere, D. 141
Quay, P. M. 206, 208, 230, 231-232, Sharp, D. 55
236, 237 Shimony,A. 28,48,49,5~56
Quine, W. V. O. 24,75,76,87,92, 158 'Simplicio' 117, 118, 119, 122, 124,
125, 131, 134, 141
Ramsey, William 6 Soal, S. G. 47
Randall,J. H. 120,141 Spencer, H. 1, 14, 24, 27
Reece, G. 54, 56 Stallo, J. B. 3, 27
Rees, M. J. 113 Stegrniiller, W. 238
Reichenbach, H. 13, 24, 27, 58, 187, Strong, J. V. 139
189 Stuart, E. B. 55
Remes, U. 116,120-121,140 Siissmann, G. 187,189,248,254
Renninger, M. 48, 56 Swanson, J. W. 92
Renouvier, C. B. 19 SzabO, A. K. 141
INDEX OF NAMES 259

Terietskii, Y. P. 187, 189 Whitrow, G. J. 113


Thirring, W. 74 Whittaker, E. T. 27
Thomson, William (Lord Kelvin) 9, 16, Wiechert, E. 187,189
18 Wiener,.P. 120,141
Tisza, L. 213, 230 Wightman, A. S. 51,56
Tomonaga, S. 45 Wigner, E. P. 35, 45, 50, 51, 55, 146,
Truesdell, C. 206 157, 158, 250, 252, 254
Windelband, W. 4
van der Deale, W. 238 Wisan, W. L. 139, 142
van der Waals, J. D. 39, 55 Witte, H. 27
Vaucouleurs, G. H. de 113 Witten, L. 113
Vigier, J.-P. 18,20,28
Voigt, W. 35
Yallop, B. D. 113
Yanase, M. M. 146, 158
Wallace, W. 116, 139, 141
Yang, C. N. 53
Wartofsky, M. W. 139
Yilniaz, H.35, 53,55
Watson, J. 235
Yourgrau, W. 28
Weinberg, S. 59,74, 112
Weizsiicker, C. F. von 253,254
Weyl, H. 10, 27 Zeno of Elea 4
Wheeler, J. A. 111 Zermelo, E. 39,52
Whitehead, A. N. 10,19,22,23,28 Zucker, F. J. 239

You might also like