Cartwright, N., How The Laws of Physics Lie 1983 PDF
Cartwright, N., How The Laws of Physics Lie 1983 PDF
Cartwright, N., How The Laws of Physics Lie 1983 PDF
OF PHYSICS LIE
NANCY CARTWRIGHT
!~
I
Mexico City Nairobi Paris Singapore
Taipei Tokyo Toronto
This book is sold subject to the condition that it shall not, by way
of trade or otherwise, be lent, re-sold, hired out or otherwise circulated
without the publisher's prior consent in any form of binding or cover
other than that in which it is published and without a similar condition
including this condition being imposed on the subsequent purchaser
"
British Library Cataloguing in Publication Data
Cartwright, Nancy
. i How the laws of physics lie.
1. Physics-Philosophy
I. Title
530'.01 QC6
ISBN 0-19-824704-4
9 10 8
Introduction
[The] velocity (with which the gas slides over the surface) and the cor-
responding tangential stress are affected by inequalities of temperature
9 See the January 1982 issue of Synthese, which is devoted to this topic,
and further references therein.
10 In correspondence in June 1981.
INTRODUCTION 9
at the surface of the solid, which give rise to a force tending to make
the gas slide along the surface from colder to hotter places. 11
Here is another, this one critical to his argument that the
vanes are not pushed by pressure normal to the surface:
When the flow of heat is steady, these forces (the total forces acting
in all directions) are in equilibrium.P
Maxwell's explanation of exactly how the motion in radio-
meters takes place will not be right unless these principles
are true. But these are not fundamental laws. Maxwell
sets his particular causal story into the framework of the
developing kinetic theory of gases. It is useful to contrast
the two specific laws quoted, about what happens in radio-
meters, with two fundamental equations from this basic
theory which Maxwell uses. In his derivation, he employs
both Boltzmann's equation
d 0
+ dz[Q(w+~-W)]=POtQ· (2)
1l James Clerk Maxwell, 'On Stresses in Rarified Gases Arising from Inequali-
Again, from Schlichting (1968) one can construct a partial list for in-
compressible viscous flow: 16
The second list gives seven more examples. But even these
fourteen examples do not provide a rigorous tie between
fundamental theory and practical circumstance. Van Dyke
concludes:
It is typical of these self-similar flows that they involve idealized
geometries far from most shapes of practical interest. To proceed
further one must usually approximate. I?
3. CONCLUSION
The picture of science that I present in these essays lacks the
purity of positivism. It is a jumble of unobservable entities,
causal processes, and phenomenological laws. But it shares
one deep positivist conviction: there is no better reality
besides the reality we have to hand. In the second sentence
of this introduction I characterized the distinction between
phenomenological and theoretical laws: phenomenological
laws are about appearances; theoretical ones about the reality
behind the appearances. That is the distinction I reject.
Richard Feynmann talks about explaining in physics as
fitting phenomena into 'the patterns of nature'. But where
are the patterns? Things happen in nature. Often they happen
in regular ways when the circumstances are similar; the
same kinds of causal processes recur; there are analogies
between what happens in some situations and what happens
in others. As Duhem suggests, what happens may even be
organized into natural kinds in a way that makes prediction
easy for us (see the last section of 'When Explanation Leads
to Inference'). But there is only what happens, and what
we say about it. Nature tends to a wild profusion, which
our thinking does not wholly confine.
The metaphysical picture that underlies these essays is
an Aristotelean belief in the richness and variety of the
concrete and particular. Things are made to look the same
only when we fail to examine them too closely. Pierre
Duhem distinguished two kinds of thinkers: the deep but
narrow minds of the French, and the broad but shallow
minds of the English. The French mind sees things in an
elegant, unified way. It takes Newton's three laws of motion
and turns them into the beautiful, abstract mathematics of
Lagrangian mechanics. The English mind, says Duhem, is
an exact contrast. It engineers bits of gears, and pulleys, and
keeps the strings from tangling up. It holds a thousand
different details all at once, without imposing much abstract
order or organization. The difference between the realist
and me is almost theologica1. The realist thinks that the
creator of the universe worked like a French mathematician.
But I think that God has the untidy mind of the English.
20 INTRODUCTION
O. INTRODUCTION
.0 Roger Rosenkrantz and Persi Diaconis first pointed out to me that the
feature of probabilities described here is called 'Simpson's Paradox', and the
reference for this example was supplied by Diaconis .
•1 See Peter J. Bickel, Eugene A. Hammel and J. William O'Connell, 'Sex Bias
in Graduate Admissions: Data from Berkeley', in William B. Fairley and Frederick
Mosteller, Statistics and Public Policy (Reading, Mass: Addison-Wesley, 1977).
38 CAUSAL LAWS AND EFFECTIVE STRATEGIES
20
.=c
iA
iA
80
80
A
A 20
12 8 42 38
8 18
CHART 1
4. CONCLUSION
25 Brian Skyrms, Causal Necessity (New Haven: Yale University Press, 1980).
Essay 2
O. INTRODUCTION
6 Miles V. Klein, Optics (New York: John Wiley and Sons, 1970), p. 21,
italics added. 8 is the angle of incidence.
THE TRUTH DOESN'T EXPLAIN MUCH 47
Refined Snell's Law: For any two media which are optically isotropic,
at an interface between dielectrics there is a refracted ray in the second
medium, lying in the plane of incidence, making an angle 0 t with the
normal, such that;
sin 0 /sin 0t = nZ/nl'
The Snell's law of page 21 in Klein's book is an example
of a ceteris paribus law, a law that holds only in special
circumstances-in this case when the media are both iso-
tropic. Klein's statement on page 21 is clearly not to be
taken literally. Charitably, we are inclined to put the modi-
fier 'ceteris paribus' in front to hedge it. But what does this
ceteris paribus modifier do? With an eye to statistical versions
of the covering law model (Hempel's I-S picture, or Salmon's
statistical relevance model, or Suppes's probabilistic model
of causation) we may suppose that the unrefined Snell's
law is not intended to be a universal law, as literally stated,
but rather some kind of statistical law. The obvious candidate
is a crude statistical law: for the most part, at an interface
between dielectric media there is a refracted ray ... But
this will not do. For most media are optically anisotropic,
and in an anisotropic medium there are two rays. I think
there are no more satisfactory alternatives. If ceteris paribus
laws are to be true laws, there are no statistical laws with
which they can generally be identified.
and
(x) (A (x) & ,Sex) 4 ,lex)).
5. CONCLUSION
O. INTRODUCTION
There is a view about laws of nature that is so deeply en-
trenched that it does not even have a name of its own. It
is the view that laws of nature describe facts about reality.
If we think that the facts described by a law obtain, or
at least that the facts that obtain are sufficiently like those
described in the law, we count the law true, or true-for-the-
nonce, until further facts are discovered. I propose to call
this doctrine the facticity view of laws. (The name is due to
John Perry.)
It is customary to take the fundamental explanatory laws
of physics as the ideal. Maxwell's equations, or Schroedinger's,
or the equations of general relativity, are paradigms, para-
digms upon which all other laws-laws of chemistry, bi-
ology, thermodynamics, or particle physics-are to be
modelled. But this assumption confutes the facticity view
of laws. For the fundamental laws of physics do not describe
true facts about reality. Rendered as descriptions of facts,
they are false; amended to be true, they lose their funda-
mental, explanatory force.
To understand this claim, it will help to contrast biology
with physics. J. J. C. Smart argues that biology has no
genuine laws of its own.! It resembles engineering. Any
general claim about a complex system, such as a radio or
a living organism, will be likely to have exceptions. The
generalizations of biology, or engineering's rules of thumb,
are not true laws because they are not exceptionless. Many
(though not Smart himself) take this to mean that biology
is a second-rate science. If this is good reasoning, it must be
physics that is the second-rate science. Not only do the
laws of physics have exceptions; unlike biological laws, they
are not even true for the most part, or approximately true.
1 See J. J. C. Smart, Philosophy and Scientific Realism (London: Routledge
and Keegan Paul, 1963).
DO THE LAWS OF PHYSICS STATE THE FACTS? 55
The gymnotoids [American knife fish] are slender fish with enormously
long anal fins, which suggest the blade of a knife of which the head
is a handle. They often swim slowly with the body straight by un-
dulating this fin. They [presumably 'always' or 'for the most part']
are found in Central and South America ... Unlike the characins
they ['usually'?] hide by day under river banks or among roots, or
even bury themselves in sand, emerging only at night."
4See Bas van Fraassen, The Scientific Image (Oxford: Clarendon Press, 1980).
5See Hilary Putnam, Meaning and the Moral Sciences (London: Routledge
and Kegan Paul, 1978) and 'Models and Reality', Journal of Symbolic Logic,
forthcoming.
DO THE LAWS OF PHYSICS STATE THE FACTS? 57
The Law of Gravitation is that two bodies exert a force between each
other which varies inversely as the square of the distance between
them, and varies directly as the product of their masses.'
I will allow that this law is a true law, or at least one that is
held true within a given theory. But it is not a very useful
law. One of the chief jobs of the law of gravity is to help
explain the forces that objects experience in various complex
circumstances. This law can explain in only very simple, or
ideal, circumstances. It can account for why the force is as
it is when just gravity is at work; but it is of no help for
cases in which both gravity and electricity matter. Once the
ceteris paribus modifier has been attached, the law of gravity
is irrelevant to the more complex and interesting situations.
This unhappy feature is characteristic of explanatory laws.
I said that the fundamental laws of physics do not represent
the facts, whereas biological laws and principles of engineer-
ing do. This statement is both too strong and too weak.
Some laws of physics do represent facts, and some laws of
biology-particularly the explanatory laws-do not. The
failure of facticity does not have so much to do with the
nature of physics, but rather with the nature of explanation.
We think that nature is governed by a small number of simple,
fundamental laws. The world is full of complex and varied
phenomena, but these are not fundamental. They arise from
the interplay of more simple processes obeying the basic
laws of nature. (Later essays will argue that even simple
isolated processes do not in general behave in the uniform
manner dictated by fundamental laws.)
This picture of how nature operates to produce the subtle
and complicated effects we see around us is reflected in the
explanations that we give: we explain complex phenomena
by reducing them to their more simple components. This is
not the only kind of explanation we give, but it is an import-
ant and central kind. I shall use the language of John Stuart
Mill, and call this explanation by composition of causes. 9
It is characteristic of explanations by composition of
9 John Stuart Mill, A System of Logic (New York: Harper and Brothers,
1893). See Book lll, Chapter VI.
DO THE LAWS OF PHYSICS STATE THE FACTS? 59
causes that the laws they employ fail to satisfy the require-
ment of facticity. The force of these explanations comes
from the presumption that the explanatory laws 'act' in
combination just as they would 'act' separately. It is critical,
then, that the laws cited have the same form, in or out of
combination. But this is impossible if the laws are to describe
the actual behaviour of objects. The actual behaviour is the
resultant of simple laws in combination. The effect that
occurs is not an effect dictated by anyone of the laws
separately. In order to be true in the composite case, the
law must describe one effect (the effect that actually happens);
but to be explanatory, it must describe another. There is
a trade-off here between truth and explanatory power.
for confirming laws, and on the other, with the use they are
put to for prediction, construction, and explanation. If laws
of nature are presumed to describe the facts, then there
are familiar, detailed philosophic stories to be told about
why a sample of facts is relevant to their confirmation,
and how they help provide knowledge and understanding
of what happens in nature. Any alternative account of what
laws of nature do and what they say must serve at least as
well; and no story I know about causal powers makes a
very good start.
Jq = fn(cxq)
Each of these is appropriate only when its cx is the only
relevant variable. For cross-effects we require laws of
the form.
The ground state of the carbon atom has five distinct energy
levels (see Figure 3.1). Physics texts commonly treat this
phenomenon sequentially, in three stages. I shall follow the
discussion of Albert Messiah in Volume II of Quantum
MechanicsP In the first stage, the ground state energy is
calculated by a central field approximation; and the single
line (a) is derived. For some purposes, it is accurate to
assume that only this level occurs. But some problems
's 'so
'D2
1s22s22p2
3P2
3p,
3
po
FIG. 3.1. The levels of the ground state of the carbon atom; (a) in the
central field approximation (Vi = V 2 = 0); (b) neglecting spin-orbit
coupling (V2 == 0); (c) including spin-orbit' coupling. (Source: Messiah,
Quantum Mechanics.)
(The real law will of course replace 'like that in the carbon
atom' by a mathematical description of the Coulomb po-
tential in carbon; and similarly for 'the three energy levels
pictured in (b )'.) The carbon atom itself provides a counter-
example to this law. It has a Coulomb potential of the right
kind; yet the five levels of (c) occur, not the three levels
of (b).
18 Ibid" p. 552.
19 Ibid" p. 552.
DO THE LAWS OF PHYSICS STATE THE FACTS? 69
6. CONCLUSION
not explain much. We could know all the true laws of nature,
and still not know how to explain composite cases. Ex-
planation must rely on something other than law.
But this view is absurd. There are not two vehicles for
explanation: laws for the rare occasions when causes occur
separately; and another secret, nameless device for when
they occur in combination. Explanations work in the same
way whether one cause is at work, or many. 'Truth Doesn't
Explain' raises perplexities about explanation by composition
of causes; and it concludes that explanation is a very peculiar
scientific activity, which commonly does not make use of
laws of nature. But scientific explanations do use laws. It
is the laws themselves that are peculiar. The lesson to be
learned is that the laws that explain by composition of
causes fail to satisfy the facticity requirement. If the laws
of physics are to explain how phenomena are brought about,
they cannot state the facts.
Essay 4
o. INTRODUCTION
The law of gravitation will illustrate what occurs in any exact science ...
Certain differential equations can be found, which hold at every in-
stant for every particle of the system . . . But there is nothing that
could be properly called 'cause' and nothing that could be properly
called 'effect' in such a system.'
1 Bertrand Russell, 'On the Notion of Cause', Mysticism and Logic (London:
1. EXPLAINING BY CAUSES
There are many interactions which may broaden an atomic line, but
the most fundamental one is the reaction of the radiation field on the
atom. That is, when an atom decays spontaneously from an excited
state radiatively, it emits a quantum of energy into the radiation
field. This radiation may be reabsorbed by the atom. The reaction
of the field on the atom gives the atom a line width and causes the
original level to be shifted as we show. This is the source of the natural
line width and the Lamb shift. 3
4 Ibid., p. 289.
5 See G. S. Agarwal, Quantum-Statistical Theories of Spontaneous Emission
and their Relation to Other Approaches (Berlin: Springer-Verlag, 1974).
6 II. Haken, 'The Semiclassical and Quantum Theory of the Laser', in S. M.
Kay and A. Maitland (008), Quantum Optics (London: Academic Press, 1970),
p.244.
80 THE REALITY OF CAUSES
approx.
1
direct approx.
solution solution ~
"C
o
;;
E
..E
c
o
.,
o
C.
0.
<I:
I
----.,
I
I I
I I
I I
I I
I I
I I neglect of higher
order derivatives
I I
_I L_
--------1
Quantum mechan. general. I
Langevin equations Fokker-Planck equation I
field + atoms field + atoms
I
I
I
I
This box I
contains the I
II exact equations JI
FIG. 4.1. Family tree of the quantum theory of the laser. (Source:
Haken, 'The Semiclassical and Quantum Theory of the Laser")
4. CONCLUSION
Perrin did not make an inference to the best explana-
tion, only an inference to the most probable cause. This
is typical of modern physics. 'Competing' theoretical
10 G. H. Harman, 'Inference to the Best Explanation', Philosophical Review
74 (1965), pp. 88-95.
86 THE REALITY OF CAUSES
O. INTRODUCTION
1. V AN FRAASSEN'S ATTACK
3. AN OBJECTION
4. CONCLUSION
O. INTRODUCTION
A long tradition distinguishes fundamental from phenomeno-
logical laws, and favours the fundamental. Fundamental laws
are true in themselves; phenomenological laws hold only on
account of more fundamental ones. This view embodies an
extreme realism about the fundamental laws of basic explana-
tory theories. Not only are they true (or would be if we had
the right ones), but they are, in a sense, more true than the
phenomenological laws that they explain. I urge just the
reverse. I do so not because the fundamental laws are about
unobservable entities and processes, but rather because of
the nature of theoretical explanation itself. As I have often
urged in earlier essays, like Pierre Duhem, I think that the
basic laws and equations of our fundamental theories organ-
ize and classify our knowledge in an elegant and efficient
manner, a manner that allows us to make very precise calcula-
tions and predictions. The great explanatory and predictive
powers of our theories lies in their fundamental laws. Never-
theless the content of our scientific knowledge is expressed
in the phenomenological laws.
Suppose that some fundamental laws are used to explain
a phenomenological law. The ultra-realist thinks that the
phenomenological law is true because of the more funda-
mental laws. One elementary account of this is that the
fundamental laws make the phenomenological laws true.
The truth of the phenomenological laws derives from the
truth of the fundamental laws in a quite literal sense-
something like a causal relation exists between them. This is
the view of the seventeenth-century mechanical philosophy
of Robert Boyle and Robert Hooke. When God wrote the
Book of Nature, he inscribed the fundamental laws of
mechanics and he laid down the initial distribution of matter
in the universe. Whatever phenomenological laws would be
true fell out as a consequence. But this is not only the view of
FOR PHENOMENOLOGICAL LAWS 101
mations Take Us Away from Theory and Towards the Truth' (unpublished manu-
script: Stanford University and Pacific Lutheran University).
104 FOR PHENOMENOLOGICAL LAWS
result for V E
given by [2tT
~ Po)II
will differ insignificantly from the result given by (6.1).
Given this insignificant variation, we can approximate VE by
using
(6.2)
I~ ) ] 1
PT - Po 1 2"
FIG.6.1
------------------1
I ale
I I
I -'----I
11
I 'b I 12
Rs- I I~
I I +
I I
+
II r, t Ie IRLV2
I
I 'c I
I
L ~ I
FIG. 6.2
1-----------------\
I, I
R s -----. rb r2 I-4--12
+
+
VJr 'n
I I
~-------------------
FIG. 6.3
9 Eugen Merzbacher, Quantum Mechanics (New York: John Wiley & Sons,
1970), pp. 484-5.
10 V. Weisskopf and E. Wigner, 'Die Rechnung der naturlichen Linienbreite
auf Grund der Diracschen Lichttheorie', Zeitschrift fiir Physik 63 (1930),
pp.54-73.
114 FOR PHENOMENOLOGICAL LAWS
f
I I
x(t) = x(to) +:- [V(t - to), x(to)] dt +
ih to
2 ,
I t,t /I I /I
+ ( ih )
Jo dt Jo dt [V(t - to), [V(t - to), x(to)] ].
We could now pull out the terms which are slowly varying
in w, and do the W integral, giving a delta function in t:
11 William H. Louisell, Quantum Statistical Properties of Radiation (New
York: John Wiley & Sons, 1973), p. 341.
116 FOR PHENOMENOLOGICAL LAWS
= g2(Weg)9&(Weg)1T f: l /
dt ceCt )5(t - t').
dCe 'Y
-=--c (t)
dt 2 e
and finally,
ceCt) = exp(-'Y/2t).
But in moving so quickly we lose the Lamb shift-a small
displacement in the energy levels discovered by Willis Lamb
and R. C. Retherford in 1947. It will pay to do the integrals
in the opposite order. In this case, we note that ce(t) is itself
slowly varying compared to the rapid oscillations from the
exponential, and so it can be factored out of the t' integral,
and the upper limit of that integral can be extended to
infinity. Notice that the extension of the t' limit is very
similar to the Markov approximation already described, and
the rationale is similar. We get
de,
- =
J +00
g2 (w)9&(w)ce(t)exp {i(weg - w)t}dw X
dt -00
oo
X Jo exp{ -i(weg - w)t'}dt'
+OO
=
J -00 g2 (w)9&(w)ce (t)exp{i(weg - w)t} dw X
FOR PHENOMENOLOGICAL LAWS 117
(.gil(X) = principal part of x.)
de,
dt =-
('Y2.+ i~WJCe(t).
.\
Here ~w is the Lamb shift. The second method, which
results in a Lamb shift as well as the line-broadening 'Y,is what
is usually now called 'the Weisskopf-Wigner' method.
We can try to be more formal and avoid approximation
altogether. The obvious way to proceed is to evaluate the
Laplace transform, which turns out to be
then
ce(t) = -.
1 f€+i~ exp(M)ce(-6)d-6.
21Tl €-i~
--=
i
M€
-'Y.
2
+ 1~W = l'Hfl
-6-->0+
"L
t weg -
get.
-
wf + 1-6
This term will give us the exponential we want:
13 See Rolf Winter, 'Large- Time Exponential Decay and "Hidden Variables" "
Physical Review 126 (1962), pp. 1152-3.
14 See D. K. Butt and A. R. Wilson, 'A Study of the Radioactive Decay Law',
3. CONCLUSION
O. INTRODUCTION
3. PHYSICS AS THEATRE
IS Ibid., p. 82.
Essay 8
O. INTRODUCTION
We saw in the last chapter that the bridge principles in a
theory like quantum mechanics are few in number, and
they deal primarily with highly fictionalized descriptions.
Why should this be so? Some work of T. S. Kuhn suggests
an answer. In his paper, 'A Function for Measurement in
the Physical Sciences', and in other papers with the word
'function' in the title, Kuhn tries something like functional
explanations of scientific practice. Anthropologists find
a people with a peculiar custom. The people themselves give
several or maybe no reasons for their custom. But the anthro-
pologist conjectures that the custom remains among these
people not only for their avowed reasons, but also because
other customs or ecological conditions make it very difficult
for their society to survive without it. This then is the 'func-
tion' of the custom in question, even though it is not practised
with any conscious awareness of that function. Naturally
all functional explanations have a dubious logic, but they
do often bring out instructive aspects of the custom in
question.
Now let us ask what function might be served by having
relatively few bridge principles to hand when we are engaged
in constructing models of phenomena. Kuhn concludes his
paper on measurement by saying he believes 'that the nine-
teenth century mathematization of physical science produced
vastly refined professional criteria for problem selection and
that it simultaneously very much increased the effectiveness
of professional verification procedures';' I think that some-
thing similar is to be said about having a rather small number
of bridge principles. The phenomena to be described are
endlessly complex. In order to pursue any collective research,
a group must be able to delimit the kinds of models that
1 T. S. Kuhn, 'A Function for Measurement in the Physical Sciences', in T. S.
Kuhn, The Essential Tension (Chicago: University of Chicago Press, 1977), p. 220.
144 THE SIMULACRUM ACCOUNT OF EXPLANATION
Mode of
radiation Reservoir I
field
Atom-Field
interaction
N three- Pumping I
level
atoms Damping I
asat = ~ [S,
ih
u; + W] +[as)
\at F
+ (as) .
at A
SO, with the situations we study, the laws which govern the
model could be presumed to apply to the real situations as
well. But models are almost never realistic in the first sense;
and I have been arguing, that is crucial to how physics works.
Different incompatible models are used for different pur-
poses; this adds, rather than detracts, from the power of the
theory. We have had many examples already but let me
quote one more text describing the variety of treatments
available for lasers:
A number of authors have treated idealized interacting systems as
models for lasers. Extensive studies have been carried out by Lax,
Scully and Lamb, Haken, Sauerman, and others. Soluble models have
been examined by Schwab le and Therring. Several simplified dynamical
models for devices of various sorts are given in the last chapter of
l..ouisell's book.!"
And so on.
There has been a lot of interest in models among philo-
sophers of science lately. It will help to compare the use I
make of models with other accounts. First, Redhead and
Cushing: both Michael Redhead 14 and James Cushing."
have recently done very nice investigations of models in
mathematical physics, particularly in quantum mechanics
and in quantum field theory. Both are primarily concerned
not with Hesse's analogical models, but with what Redhead
calls 'theoretical models', or incomplete theories (Cushing's
modelj-e-guinea-pig or tinker toy models). Although, like
me, Cushing explicitly says that models serve to 'embed'
an account of the phenomena into a mathematical theory,
he and Redhead concentrate on a special kind of model-a
theory which is admittedly incomplete or inaccurate. I am
concerned with a more general sense of the word 'model'.
I think that a model-a specially prepared, usually fictional
description of the system under study-is employed when-
ever a mathematical theory is applied to reality, and I use
the word 'model' deliberately to suggest the failure of exact
we will find that it is very easy to treat the problem by the use of differ-
ential equations. This method obscures the physical origin of the index
(as coming from the re-radiated waves interfering with the original
waves), but it makes the theory for dense materials much simpler.i"
But what is it for a theoretical treatment to 'tell' a causal
story? How does Feynman's study of light in Volume I
'make clear' the physical principles that produce refraction?
I do not have an answer. I can tell you what Feynman
does in Volume I, and it will be obvious that he succeeds in
extracting a causal account from his model for low density
materials. But I do not have a philosophic theory about
how it is done. The emphasis on getting the causal story
right is new for philosophers of science; and our old theories
of explanation are not well-adapted to the job. We need a
theory of explanation which shows the relationship between
causal processes and the fundamental laws we use to study
them, and neither my simulacrum account nor the tradi-
tional covering-law account are of much help.
Causal stories are not the only problem. Even if we want
to derive only pure Humean facts of association, the D-N
account will not do. We have seen two ways in which it
fails in earlier chapters. First, from Essay 6, the fundamental
laws which start a theoretical treatment frequently get
corrected during the course of the derivation. Secondly,
many treatments piece together laws from different theories
and from different domains, in a way that also fails to be
deductive. This is the theme of Essay 3.
These are problems for any of our existing theories of expla-
nation, and nothing I have said about simulacra helps to solve
them. Simulacra do a different job. In general, nature does
not prepare situations to fit the kinds of mathematical theories
we hanker for. We construct both the theories and the objects
to which they apply, then match them piecemeal onto real
situations, deriving-sometimes with great precision-a bit
of what happens, but generally not getting all the facts straight
at once. The fundamental laws do not govern reality. What they
govern has only the appearance of reality and the appearance
is far tidier and more readily regimented than reality itself.
20 Richard Feynman, The Feynman Lecture on Physics, vol. II (Reading,
Mass: Addison-Wesley, 1964), p. 32.1.
Essay 9
I
H<,C/C~C/Br
II
c
I
c
H/ """C~ <,Br
(b) 1
FIG 9.1.0rthodibromobenzene (Source:Feynman, Lectures on Physics.)
So
It follows that
C3: Prob(y) = Prob{ (y & Sl) V (y & S2)}
Q2: t = Probts.),
and similarly for S2, we see that the classical calculation C
for the probability oflanding aty, and the quantum mechanical
calculation Q do not give the same results. They differ
by the interference terms t¢l(Y)¢i'(Y) + t¢r(y)¢2(y). The
C calculation is the one that supposes that the electrons have
a definite location at the screen and its consequences are
not borne out in experiment.
There are various ways to avoid the conclusion C4, well-
rehearsed by now. The first insists that quantum mechanical
propositions have a funny logic. In particular they do not
THE MEASUREMENT PROBLEM 177
obey the distributive law. This blocks the argument at the
move between C1 and ~. This solution was for a while
persuasively urged by Hiliary Putnarn.P Putnam has now
given it up because of various reflections on the meaning
and use of the connective or; but I am most impressed by
the objections of Michael Gardner!" and Peter Gibbins!"
that the usual quantum logics do not in fact give the right
results for the two-slit experiment.
The second well-known place to block the C derivation
is between the third and the fourth steps. The first attempt
maintained that quantum propositions have a peculiar logic;
this attempt insists that they have a peculiar probability
structure. It is a well-known fact about quantum mechanics
that the theory does not define joint probabilities for non-
commuting (i.e. incompatible) quantities. Position and
momentum are the conventional examples. But the different
positions of a single system at two different times are also
incompatible quantities, and so their joint probabilities are
also not defined. (The incompatibility of 'r at t' and 'r at t"
is often proved on the way to deriving the theorem of Ehren-
fest that, in the mean, quantum mechanical processes obey
the classical equations of motion.) So Prob(y & S1) and
Prob(y & S2) from C4 do not exist. If Prob(y) is to be cal-
culated it must be calculated in some different way-
notably, as in the Q derivation.
What does it mean in a derivation like this to deny a
joint probability? After all, by the time of C3 we have gone so
far as to admit that each electron, individually, has a well-
defined position at both the screen and at the plate. What
goes wrong when we try to assign a probability to this con-
junctive event? Operationally, the failure of joint distributions
must show up like this. We may imagine charting finite
histograms for the joint values of 'r at t' and 'r at t", but
the histograms forever bounce about and never converge to
a single curve. There is by hypothesis some joint frequency
Plate
-----~~~---__+---------z
Plate
----------==--''-------l------_z
the detector across the plate (Figure 9.3). But the same
effect is achieved by fixing the location of the detector
and varying the centre of the wave (Figure 9.4). Looked
at in this way, the consistency result can be seen as a trivial
consequence of a fundamental theorem of scattering theory.
This theorem states that the scattering cross section, both
total and differential, is a constant independent of the
shape or location of the incoming wave packet. The total
scattering cross section is essentially the ratio of the total
probability for the particle to be scattered, divided by the
probability for crossing the detector. Roughly, the theorem
assures that the probability of scattering from the detector
for a wave centred at fO, divided by the probability for
'being at' the detector, is a constant independent of fo.
This is just what is required by (9.2).
There is one difficulty. Standard textbook proofs of
this theorem do not establish the result for arbitrary waves,
but only for wave packets that have a narrow spread in
the direction of the momentum. This is not enough to
186 THE MEASUREMENT PROBLEM
z z z
x y x y x y
t =0 t='/.ll/I1E t=%1i/tJ.E
z z
x y x y
t=%1i/I1E t=f1/I1E
This decays at the same rate whether on the wire on which it is origin-
ally deposited, or in the solution of hydrochloric or nitric acid. The
excited radioactivity produced by the radium emanation appears
analogous."
The third fact is the one of relevance to us. From Ruther-
ford and Soddy's introduction: 'Radioactivity is shown to be
accompanied by chemical changes in which new types of
matter are being continuously produced.F" During decay,
uranium 238 is transformed into thorium 234. When the
alpha particle is emitted, the state of the material changes.
This is exactly analogous to Einstein's treatment of atomic
decay in his derivation of the Plank law of black body
radiation (though Einstein's real feeling about this situation
seems to have been far more ambiguous). Concerning what
we now call spontaneous emission he says, 'This is a transi-
tion from the state Zm into the state Z; together with
emission of the radiation energy Em - En. This transition
takes place without any external influence. One can scarcely
avoid thinking of it as like a kind of radioactive reaction. '29
Bohr too has the same picture when he first quantizes the
energy levels of the atom. Only certain orbits are allowed
to the electrons. In radiation the electron changes from one
fixed orbit to another, emitting a quantum of light energy.
In discussing the spectrum of hydrogen Bohr says, 'During
the emission of the radiation the system may be regarded
as passing from one state to another.P? On the old-quantum-
theory picture the time of decay is undetermined. But when
decay occurs, a photon, or an alpha particle or a beta particle
is given off, and the emitting material changes its state.
Contrast this with the new-Quantum-theory story. This is
the story we read from the formalism of the developed
mathematical theory. On this story nothing happens. In
I should mention that Davies himself does not use the formal-
ism in the way that I urge; for he is at pains to embed the
non-unitary evolutions he studies into a Schroedinger evolu-
tion on a larger system. This goes along with his suggestion
that non-unitarity is a mark of an open system, one which
is in interaction with another. Open systems are presumably
parts of larger closed systems, and these in Davies's account
always undergo unitary change. I think this view is mistaken,
for the reasons I have been urging throughout this essay.
If the wave packet is not reduced on the larger system, it
is not in fact reduced on the smaller either. The behaviour
of the smaller system at best looks as if a reduction has
occurred, and that is not good enough to account for
measurements or for preparations.
I have been urging that, if the quantum statistical pro-
gramme can work, the measurement problem becomes a
psuedo-problem. But other, related, problems remain. These
are the problems of how to pick the right operators, unitary
or no, to represent a given physical situation. This is the
piece by piece work of everyday physics, and it is good
to have our philosophical attentions focused on it again.
This is what on-going physics is about, and it knows no
general procedure. In quantum mecahnics the correspond-
ence principle tells us to work by analogy with classical
mechanics, but the helpfulness of this suggestion soon runs
out. We carry on by using our physical intuitions, analogies
we see with other cases, specializations of more general
considerations, and so forth. Sometimes we even choose
the models we do because the functions we write down are
ones we can solve. As Merzbacher remarks about the Schroe-
dinger equation:
Quantum dynamics contains no general prescription for the con-
struction of the operator H whose existence it asserts. The Hamil-
tonian operator must be found on the basis of experience, using the
clues provided by the classical description, if one is available. Physical
insight is required to make a judicious choice of the operators to be
used in the description of the system ... and to construct the Hamil-
tonian in terms of these variables. 39
44 Ibid., p. 505.
THE MEASUREMENT PROBLEM 209
21Th2 A I
Tkk, = -3 fdk)
J.l.L
we get
4 3
12 81T I 12
,to
I <1/Jkll/)k) = I + k3L6 A(O) + 41T
kL3 ImA(O). (2)
and
E( 4.8)b: R ~ t(WIIi)
Using E(4.8), R 2 + j2 = (L WiRi)2 + (L Wi.!;)2 • But in general
L -d (VJk,IUIVJk)(VJk IUIVJk')
k'Ed!1 dt
= lim _1
~ /'->0 At
(L
k'Ed!1
(VJk' IU(t + At) IVJk)( VJk IU(t + At)IVJk')
I W; Wi'
But
LI
" (!k(O)Ojk) =L
2"· w;(fJ(O)) 2.
idealization 48, 109 ff., 134, 136, 147, open systems 198,205
148,153,155,158 optical theorem 206 ff,
inductive-statistical model of explana- order of integration 120 ff.
tion 26, 30,44,47
inference to best explanation 4, 6, partial conditional probability 42
15,75,82,83,85,87-9 partitioning 36 ff., 41, 43
inference to most likely cause 6, 85, Pauli equation 114, 115, 124
92,94 phenomenological laws 1 ff., 8,11,16,
infinite potentials 150, 153 19, 85, 88 rr., 93 ff., 100 rr.,
interactions, causal 30-2, 63 106, 160, 161
see also composition of causes phenomenological models (phenomeno-
and cross-effects logical terms) 148,151
internal principles 131, 132, 135, 139 photographic plate 182 ff.
positions of quantum systems 163,
joint effects 43 166, 167, 171, 178, 182, 183,
see also spurious cause 186,188 ff.
joint probability distribution 156, 177, position probabilities 164, 174, 179,
178,181 180,188,190,191
preparation process (problem of prep-
Lamb shift 116, 117, 137, 138 aration) 174, 194, 205, 215,
in the excited state 119-23 216
in the ground state 119, 123-7 prepared descriptions (unprepared de-
lasers 3, 79, 80, 98, 128 ff., 145; 146, scriptions) 15, 17, 133, 139,
148, 149, 152, 153, 157 ff., 147,160,162
161,186,189,198,204 probability distributions, classical 153
laws of association 10, 21, 25, 26, 28, ff.
35,40 ff., 46 probabilistic model of causation 44, 47
line broadening 77 ff. projection postulate 162, 167, 170,
198
macroscopic observable (macroscopic property of convenience 153, 156
system) 167, 168, 170 ff., 195,
196 quantum electrodynamics (quantum
Markov approximation 113 ff., 123, field theory) 7, 8
124, 133, 146, 189, 197 quantum logic 164, 177, 181
master equation 79, 81, 113, 114, quantum statistical mechanics 164,
123,125,145,197 199 ff., 204, 205
measurement problem 18, 162-5
mechanical philosophy 100, 107 Rabi-flopping 196, 199
Mill's methods 7,98 radioactivity 25,27, 28, 78, 93, 118,
mixtures 168, 170, 171, 194, 204, 192,193
206,212,215 see also decay, exponential
models, explanatory 4, 11, 12, 15, radiometer 5, 6, 8, 9, 11, 81, 82,
17, 44, 83 ff., 95, 104, 107, 154 ff.
111, 129, 140, 143, 144, 148, rate equations 189, 190
152,154,156 ff., 197, 198,202 realistic in the first sense (in the
second sense) 150, 152
Newton's laws 63 reduction of the wave packet 163, 164,
167,170,174,194 ff., 206-16,
observable (unobservable) 1, 2, 5, 6, appendix
19,56,83,100,106,159,160 redundancy 76, 79, 90
Ohm's law 63 reference class problem 28, 29
old quantum theory (new quantum reservoir 113 rr., 129, 133, 146, 148
theory) 137, 192 ff. ff., 189, 197, 202
SUBJECT INDEX 221