Vous êtes sur la page 1sur 15

Stud. Hist. Phil. Mod. Phys., Vol. 30, No. 2, pp.

267 } 281, 1999


 1999 Elsevier Science Ltd. All rights reserved
Printed in Great Britain
1355-2198/99 $ - see front matter

E SS AY R E V I E W
The Phenomenological Approach to Physics
Steven French*
Kostas Gavroglu, Fritz ondon: A Scienti,c Biography (Cambridge: Cambridge
University Press, 1995), xxiv#299 pp. ISBN 0521432731.

1. Introduction
Kuhn famously described &normal' science as &mop-up work' (Kuhn, 1970, p. 24).
The problem for the historian is how to make this interesting and engaging. One
solution is to choose a &key' "gure who lies close to the establishment of
a particular paradigm. Thus Klein gave us Ehrenfest, remembered not as one of
the founders of quantum mechanics but as a brilliant teacher and expositor, who
helped to lay bare the foundations of the new framework (Klein, 1970). Gavroglu in turn gives us Fritz London, a &key' "gure in the further development of
the new quantum mechanics and the means by which Gavroglu can explore the
rich structure of post-revolutionary modern physics. The comparison with
Klein's masterpiece is sustainable, I believe. Like Klein, Gavroglu not only
presents, in a clear and accessible manner, the scienti"c &product' of the life, but
also gives us a feel for the life itself. And this life was, if not &normal', then not
atypical for a gifted Jewish scientist of the time.
Born in Breslau at the turn of the century, Fritz London became interested in
fundamental problems of epistemology while still at school but decided to study
physics with Sommerfeld at Munich. While there he came into contact with the
philosopher PfaK nder, a well-known phenomenologist, who was so impressed
with one of London's philosophical pieces that he o!ered to accept a rewritten

(Received 20 July 1998)


* Division of History and Philosophy of Science, School of Philosophy, University of Leeds,
Leeds, LS8 2LP, U.K. (e-mail: s.r.d.french@leeds.ac.uk).
PII: S 1 3 5 5 - 2 1 9 8 ( 9 9 ) 0 0 0 0 6 - 4

267

268

Studies in History and Philosophy of Modern Physics

version as a thesis for the philosophy degree. After graduation, London's interest
in epistemological problems in physics led him to GoK ttingen, where Born
insisted that he do some &straight' physics. Subsequently his scienti"c career
took o! and the list of achievements includes: the quantum mechanical explanation of the homopolar bond with Heitler, which laid the foundations of quantum
chemistry; the careful work on molecular forces, understood through the uncertainty principle; the famous theory of superconductivity with his brother, which
took account of the Meissner e!ect; the explanation of super#uidity in terms of
Bose}Einstein condensation; and the monograph with Bauer on the measurement problem, well known to students in the philosophy of physics. He #ed
Germany for Britain in 1933 but, after failing to secure a permanent position,
moved to Paris before emigrating to the U.S.A., where he was o!ered a job in the
Chemistry Department at Duke University. While he was there his wonderful
monograph Super-uids was published and in 1953 London was awarded the
Lorentz Medal. In 1954 he died of a heart condition.
These are the bare bones of a life that, at least for that period, might be
described as normal. Gavroglu gives us the #esh too, from Fritz's relationship
with his brother Heinz to the acrimonious dispute with von Laue over the
latter's contribution to the theory of superconductivity. The science is handled
deftly and the life regarded sympathetically. With an afterword from Bardeen,
written the year before his death, the book sits comfortably next to Klein's.
But of course, London was also abnormal, as Gavroglu emphasises: London
never abandoned his philosophical roots in phenomenology. There is an interesting book to be written on the inter-relationship between science and philosophy in this period, but it is here that Gavroglu's touch is less sure and his
analysis less insightful. The problem lies in getting to grips with phenomenology
itself. A little over four pages is just not su$cient to cover the work of PfaK nder,
Becker and Husserl, and although over seven pages are devoted to London's
philosophy thesis, much of it is opaque, if not incomprehensible. However,
a number of crucial points are discernible, the most important being a nonreductionist, holistic view of scienti"c theories. This &macrological' view of
theories as structured wholes &would become one of London's reference points in
developing his own ideas about deductive systems' (p. 13). Here Gavroglu leans
heavily on Mormann who argues that Husserl's work can be seen as a forerunner of the semantic approach to theories (Mormann, 1991). Whether or not this
is the case it is surely * dare I say it to a historian * unbearably whiggish to
write,
[2] even before starting his thesis, London was quite sensitive and receptive to
ideas related to the semantic approach in the philosophy of science (p. 15).

London also had a two-stage view of the relationship between theory and
reality: the "rst stage takes us from reality to experience, as the phenomenon in
 Page references are to Gavroglu's Fritz ondon: A Scienti,c Biography (1995), unless otherwise
indicated.

The Phenomenological Approach to Physics

269

question is appropriately formulated and in the second we move from experience to language and an explanatory schema. Gavroglu insists that such distinctions are crucial to understanding London's work and it is interesting to
consider in a little more detail how this &normal' scientist's philosophy of science
a!ected his scienti"c work.

2. Eating Chemistry with a Spoon: the Group-theoretical &Reduction' of


Chemistry to Quantum Mechanics
As Heitler subsequently expressed it, the &fundamental problem' of chemistry
was &the reason why atoms prefer to combine and form molecules consisting of
a de,nite number of atoms' (Heitler, 1967, p. 13). His reductionist tendency is
evident in his insistence that this was a physical problem, with the attraction
explained in terms of &physical laws prevailing within the atom' (ibid.) without
the necessity of introducing further chemical forces. However, it became evident
that the attraction between two hydrogen atoms could not be accounted for in
terms of Coulomb forces; the key, as Heitler realised one hot and disagreeable
day in ZuK rich (p. 45), lay with the exchange integral, previously introduced by
Heisenberg. This was something purely quantum mechanical, with no classical
analogue and as Gavroglu notes, it is this focus on purely quantum mechanical
explanations that also marks London's work.
Carson has nicely charted the development of the exchange concept and, in
particular, the shift from the idea of a literal exchange of electrons to the
understanding of its basis in the non-classical indistinguishability of the electrons (Carson, 1996). The Heitler}London paper is situated right at the heart of
contemporary struggles to grasp the essentially metaphysical implications of the
new quantum mechanics. On the one hand, the mathematics itself harkened
back to the literal picture of exchange and the idea of beats; on the other, Heitler
and London noted that the electrons could not be regarded as labelled. On this
fundamental basis the electronic wave function of the two-atom system could be
written in either symmetric or anti-symmetric form. With the electron spins
incorporated, the Pauli Exclusion Principle dictates that the anti-symmetric
form be chosen, with spins anti-parallel. This corresponds to the state of lower
energy and hence attraction. With the introduction of spin and the Exclusion
Principle chemical valence and saturation could be understood and the &problem of chemistry' solved.
London subsequently incorporated spin into his general formulation of the
Exclusion Principle suitable for cases with more than two electrons and this
proved more convenient for his later work with group theory (p. 49). The
group-theoretical approach to these problems in chemistry was initiated by
Heitler, after moving to GoK ttingen * where although Born was negative (p. 56)
 In a letter to Born in 1936, London wrote: &[2] we were very proud when we realized that we get
the exchange degeneracy because of the similarity of the electrons' (pp. 87}88; London's emphasis).

270

Studies in History and Philosophy of Modern Physics

Heitler learned from Wigner and Weyl the recent mathematico-physical theory.
His central idea was that the Heitler}London treatment of the hydrogen bond
could be extended to other molecules if it were underpinned by the theory of
irreducible representations of the permutation group, as applied to atomic
physics by Wigner and Weyl. In this manner, he claimed, &[w]e can [2] eat
Chemistry with a spoon' (p. 54).
London, although less reductionistic than Heitler (pp. 91}92), agreed that
group theory provided a way of dealing with the many-body problem. In a series
of papers in 1928 he showed that the group-theoretic formulation of quantum
mechanics could recover the same valence numbers and satisfy the same formal
combination rules as were expressed in the chemist's semi-empirical framework
(p. 57). With this work, taken together with Heitler and Rumer's 1931 analysis of
the valence structures of polyatomic molecules, it did indeed seem that chemistry had been eaten with a spoon.
Not surprisingly, perhaps, the chemists disagreed and Gavroglu ably tracks
the ins and outs of Heitler and London's disputes with Mulliken, Pauling and
Slater (pp. 82}92). However, it is not so clear where London's philosophy "ts.
As Gavroglu himself notes: &It is ironic, that this "rst important paper of his was
such a pronounced deviation from his grand schema to view theories as wholes!'
(p. 92). Gavroglu does claim that London's subsequent group theoretical work
contains suggestions of an alternative non-reductive approach (p. 74 and pp.
91}92) but these suggestions are never made explicit. It is noted, however, that
London took the fundamental problem to be
[2] the mysterious order of clear lawfulness, which is the basis for the immense
factual knowledge of chemistry and which has been expressed symbolically in the
language of chemical formulas (quoted on p. 56).

Here a form of the two-stage process can be reconstructed, with the "rst
having already been completed by the chemists in a semi-empirical fashion and
the second providing a sound theoretical basis for the chemists' rules and
formulas. However the "rst stage notion of saturated valences could not be
explained by the &standard' principles of atomic physics, but required spin, the
Exclusion Principle and, ultimately, group theory (pp. 56}57). Perhaps in
a weak sense there is a kind of anti-reductionism here.
This is, of course, very much a hot topic among philosophers of chemistry.
Some of the issues involved are manifested in London's work through the
tension between his emphasis on purely quantum mechanical explanations and
his non-reductive philosophy. One way of easing this tension, perhaps, is to give
up the reductionist talk of di!erent &levels' and instead accept that there is one
 On the physics side, Slater was hailed as having &slain the Gruppenpest' by Condon and Shortley
in their 1935 heory of Atomic Spectra. However, their linear operators of angular momenta are
simply the generators of the rotation group SO(3).
 See, for example, Scerri (1997).

The Phenomenological Approach to Physics

271

world which is very abstractly, and perhaps only provisionally, described in


quantum mechanical terms. Because of the complexity of the phenomena, exact
derivations are simply not feasible. Shifting from the dynamics to the symmetries, as described by group theory, gives us an explanation of the phenomena, or
at least as much of one as we can have within these constraints. As Pauling
himself came to acknowledge, spin is central to our understanding of the
chemical bond and spin is purely quantum mechanical. But this is not to say
that chemists must become physicists, or the latter mathematicians. The idealised and partial nature of these explanations suggests that a degree of autonomy
will always remain. If reduction is tied to deduction from SchroK dinger's equation, then London's use of group theory is non-reductive. But this does not
imply complete theoretical autonomy. It is surely time to abandon the positivistic view of reduction as involving a distinctive entailment relation and instead
focus on alternative frameworks. Perhaps Heitler's and London's group theoretical work will repay careful study in this regard.
The in#uence of London's philosophy on his science is much more apparent
in his second major piece of work, his account of superconductivity.

3. The Superconductor as a Single Big Diamagnetic Atom


In 1928 Bloch published a quantum mechanical theory of electron conductivity which accounted for metals, semiconductors and insulators. It failed to
account for superconductivity, however, which was conceived of in terms of an
analogy with ferromagnetism, in the sense that persistent currents existing below
the critical temperature were taken to be analogous to permanent magnetisation
below the Curie point (pp. 112}113). This failure led to &Bloch's theorem':
superconductivity is impossible! It was the discovery of the Meissner e!ect in
1933 which indicated that it was the analogy which was at fault. Instead,
superconductors should be regarded as diamagnetic, a suggestion which had
been made previously by Frenkel in 1933 but which was set at the very heart of
London's account with his brother Heinz (London and London, 1935).
Gavroglu follows convention in describing the London}London model as
&phenomenological'. However, London himself rejected the term and insisted
that the model should be described as &macroscopical' since it goes beyond the
strictly phenomenological data. Furthermore, to take the model to be purely
&phenomenological' would be to ignore the importance of London's own

 Hartree, for example, was famously less than enthusiastic about the introduction of group theory,
although he concluded: &Is it really going to be necessary for the physicist and chemist of the future to
know group theory? I am beginning to think it may be' (quoted on p. 56).
 Ramsey has recently suggested the abandonment of the &levels' view of reduction in favour of
a &perspectives' approach which insists that &[2] there is only one reality; we make various
approximations to capture the properties of that reality' (1997, p. 246).
 See also Cartwright, Suarez and Shomar (1996).

272

Studies in History and Philosophy of Modern Physics

philosophical view of theoretical construction. As Gavroglu reminds us (pp.


127}128), the phenomena must "rst be formulated in a particular fashion and
then embedded in an explanatory framework. It was precisely because the
phenomena of superconductivity had been represented inappropriately to begin
with that it was concluded that it was theoretically impossible.
Thus London wrote in 1935:
The progress I claim is mainly a logical one: by a new and more cautious
interpretation of the facts I tried to avoid a fundamental di$culty (the so-called
theorem of Bloch) which stood in the way of explaining superconductivity by the
customary theory of electrons in metals and which could not be overcome as long
as one has considered this phenomenon as a limiting case of ordinary conductivity
(quoted on p. 129).

The di$culty was avoided by switching analogies, under the pressure of experimental results, and reformulating the phenomena. But this is only the "rst stage;
complete understanding can only be reached by proceeding further and arriving
at a new explanatory schema. Stopping at the "rst stage would be to leave the
process incomplete.
In the move to the second stage the &macroscopical' theory is crucial. The term
&macroscopic' had a dual meaning: "rst of all, the theory was &macroscopic' in
the straightforward sense that it was concerned with electric and magnetic "eld
strengths and the like. Secondly, and more fundamentally, it was macroscopic in
the sense that superconductivity was seen by London as a uniquely quantum
mechanical phenomenon of long-range, or macroscopic, order (the development
of this notion of macroscopic is clearly set out by Gavroglu on p. 144). It is this
latter understanding which constrains the &microphysical' or &molecular-kinetic'
explanation of the phenomenon. Thus London and London write that this &new
description [2] seems to provide an entirely new point of view for a theoretical
explanation' (p. 71). And in his 1935 Royal Society presentation (London, 1935),
Fritz London provides a &sketch' of such an explanation, elaborating on the
concluding remarks of the joint paper where a structural similarity is drawn
with Gordon's equation for electric current and charge in his relativistic formulation of SchroK dinger's theory and the suggestion is made for the "rst time that
the electrons are coupled in some way. This &microphysical' programme is &set'
by the diamagnetic analogy, thus linking these two senses of &macroscopic' (for
further details see French and Ladyman, 1997). As Gavroglu notes, this programme subsequently became a &valuable heuristic' for the work of Bardeen and
the idea of coupled electrons came to be expressed in the concept of &Cooper
pairs' (p. 209).
Thus London's second stage of understanding is already partially present in
the joint 1935 paper. The &macroscopical interpretation' constrains * via the
structural similarity with Gordon's equation and driven by the diamagnetic
analogy * the microscopical. This e!ects only a &reduction' of the possible
mechanism responsible (London, 1937, p. 795; cf. London, 1950, p. 4) and hence
the second stage remains incomplete. The macroscopic model can be seen as

The Phenomenological Approach to Physics

273

a kind of half-way house, going beyond the &phenomenological' and extending


into the &theoretical' but leaving the latter open to further elaboration and
development (cf. Gavroglu's comments on p. 143). This seems to be the view
London took in Volume II of Super-uids, subtitled &Macroscopic Theory of
Liquid Helium'. In the preface he writes:
The term &Macroscopic Theory' in the subtitle is not entirely adequate. It is meant
to indicate that, as in Volume I [which covered superconductivity], it is not the
plan to present a molecular-kinetic theory of this subject, a theory which in fact has
not yet been established. On the other hand, it did not appear reasonable to keep
strictly within the limits of phenomenological physics (1954, p. xi).

However, Gavroglu argues that there is one signi"cant di!erence between the
superconductivity and liquid helium cases which concerns the role and place of
a &molecular-kinetic' theory.

4. Super6uidity and Structures


In his last year at Oxford London became interested in what he later called
the &mystery' of the &liquid degeneracy' of liquid helium (pp. 147}148). The
demonstration that liquid helium could only be solidi"ed under pressure led to
the suggestion that the liquid passed into some kind of ordered state below the
transition temperature. London originally suggested that it had a diamond
lattice structure but this idea was subsequently abandoned in favour of a new
model in terms of Bose}Einstein condensation (pp. 152}157).
Thus in his 1938 note to Nature, London writes that &[2] in the course of
time the degeneracy of the Bose}Einstein gas has rather got the reputation of
having only a purely imaginary existence' (1938a, p. 644). After indicating that
a static spatial model of liquid helium was not possible, he went on to &imagine'
an analogy with Bloch's model of electron conductivity in metals, with the
crucial di!erence that, with helium atoms instead of electrons, &[2] we are
obliged to apply Bose}Einstein statistics instead of Fermi statistics' (ibid.,
p. 644). London then indicates that the Bose}Einstein condensation represents
a discontinuity of the derivative of the speci"c heat and although this is a phase
transition of third order, rather than one of second order as in the case of helium,
the experimental values for the transition temperature and entropy agree quite
favourably with those calculated for an ideal Bose}Einstein gas.
London expanded on this suggestion in a paper for the Physical Review
where he presented a proof of the Bose}Einstein condensation and shows how
this &highly idealized' model can give an account of the peculiar transport
phenomena of liquid helium * such as the &fountain e!ect' * to qualitative
agreement with experiment (1938b, pp. 953}954). Indeed, in Super-uids he
suggested that the above numerical agreement &would perhaps not have deserved much attention' (1954, p. 59) had not his model o!ered the promise of
a qualitative interpretation of the &super' properties and &striking peculiarities'.

274

Studies in History and Philosophy of Modern Physics

In this second 1938 paper he also considers the meaning of the condensation
and argues that it cannot represent a physical condensation in ordinary space,
since the particles do not disappear mysteriously from space. Rather: &If one likes
analogies, one may say that there is actually a condensation, but only in momentum space [2]' (1938b, p. 951, London's emphasis; cf. p. 39 of Super-uids,
Vol. II, where this is seen as a &manifestation of quantum-mechanical complementarity'). This condensation does lead to a &characteristic peculiarity'
(1938b, p. 951) in ordinary space, represented by a &peculiar omnipresence' in the
total volume available to the molecules. This is once again a &macroscopic'
quantum e!ect and London draws the comparison with superconductivity.
However, Gavroglu insists that in the case of super#uids, this &macroscopic'
e!ect was implied by the formalism of the theory itself, since it incorporated
Bose}Einstein statistics, whereas in that of superconductivity it was an interpretation imposed on the formalism. Hence, he claims, the London and London
theory could be accepted and applied without any corresponding ontological
commitment (p. 236). However, this &dramatically di!erent situation' (ibid.)
raises a problem as it seems that, in the case of liquid helium, London deviated
from his own philosophy of science, thus undermining one of the central claims
of the whole book. Can we reconstruct a two-stage process in this case?
In his note for Nature London explicitly states that his intention is to show
that a static spatial model &of whatever regular con"guration' is not possible and
then &[2] to direct attention to an entirely di!erent interpretation of this
strange phenomenon' (1938a, p. 643). This di!erent interpretation is that of
a highly non-standard liquid, one that is very similar to a gas. As in the
superconductivity case, the elaboration of this new model is motivated by
a combination of experimental and theoretical considerations in which the role
of analogy is fundamental. This is clearly set out in the "rst two chapters of
Vol. II of Super-uids (cf. Gavroglu, p. 150), where he writes,
[2] this system does not represent a liquid in the ordinary sense. There are no
potential barriers as in ordinary liquids to be overcome when an external stress is
applied. The zero point energy is so large that it can carry the atoms over the
barriers without requiring the intervention of the thermal motion. In this respect
there seems to be a greater similarity to a gas than we are used to assume in
ordinary liquids. This view is supported by the extremely signi"cant fact that liquid
helium I, which on "rst sight appears to be quite an ordinary viscous liquid,
actually has a viscosity of a type ordinarily found only in gases and not in liquids
(1954, p. 37; London's emphasis).

Having reconceptualised the phenomenon, one can then move to the second
stage. One of the characteristic features of this strange gas-like liquid is the
 It is noticeable that he again points out that in the latter case the macroscopic phenomena can be
understood in terms of a &peculiar coupling' in momentum space, &as if there were something like
a condensed phase& in this space (London, 1983b, p. 952; Gavroglu, 1995, p. 158). This suggestion
was subsequently and independently taken up by Feynman, Ginzburg and Schafroth (p. 246).
Feynman went on to develop the quantum statistical explanation of super#uidity (Mehra, 1994,
pp. 348}391).

The Phenomenological Approach to Physics

275

discontinuous transition noted above and here we have a structural similarity


with a Bose}Einstein (but not Fermi}Dirac) gas which exhibits a similar
discontinuity. On this basis the theoretical understanding can be built up. This is
explicitly the way in which London develops his approach in Super-uids and
one is tempted to suggest that here he might be reconstructing these developments in accordance with his philosophy. Of course the mechanism behind the
quantum condensation is better developed than in the superconductivity case,
but it should be noticed that this model is also both partial and, ontologically,
relatively autonomous.
It is partial because, as London emphasises, there was no adequate theory of
liquids:
[2] it is obvious that the theoretical basis given thus far is not to be considered
more than a quite rough and preliminary approach to the problem of liquid
helium, limited chie#y by the lack of a satisfactory molecular theory of liquids
(1938b, p. 954).

As for autonomy, just as the London}London model of superconductivity


could be applied and developed in the absence of a detailed mechanism for the
electron coupling, so could the &macroscopic' theory of liquid helium. After
presenting his model of a Bose}Einstein liquid, London writes that
[2] an understanding of a great number of the most striking peculiarities of liquid
helium can be achieved, without entering into any discussion of details of molecular mechanics, merely on the hypothesis that some of the general features of the
degenerating ideal Bose}Einstein gas remain intact, at least qualitatively, for this
liquid, which has such an extremely open structure. This is an assumption which
may be judged by the success of its consequences in describing the facts, pending an
ultimate justi"cation by the principles of quantum mechanics (1954, pp. 59}60; his
emphasis).

He then acknowledges Tisza as being the "rst to recognise &[2] the possibility
of evading the pitfalls of a rigorous molecular-kinetic theory by employing the
qualitative properties of a degenerating Bose}Einstein gas to develop a consistent macroscopic theory' (ibid; the changing nature of London's view of Tisza's
work is nicely documented by Gavroglu, pp. 159}163; see also pp. 198}206 and
pp. 214}217). Again, this &macroscopic' theory not only described the &facts'
concerning liquid helium II in terms of a common theoretical basis but also
made speci"c predictions of the properties of the liquid.
Viewed this way, a degree of commonality can be discerned in both cases:
there was a critical reconceptualisation of the phenomenon * from an analogy
with ferromagnetism to that with diamagnetism in the case of superconductivity,
and from a quasi-crystalline state to a Bose}Einstein liquid in the case of
Helium II * which is embedded within, or represented by means of, a &macroscopic' theory, accompanied by a move to a &molecular-kinetic' understanding
which is only partial. And in both cases, London remained convinced that the
phenomena could only be explained through the crucial role of purely quantum
mechanical notions, as Gavroglu emphasises (pp. 235}236).

276

Studies in History and Philosophy of Modern Physics

5. A Phenomenological Solution to the Measurement Problem? 9


London's only contribution to the foundations of quantum mechanics was
the monograph he wrote with Bauer on the theory of observation (London and
Bauer, 1939). The intention was to re-present, in more accessible form, von
Neumann's classic analysis and to make explicit the role of consciousness in
e!ecting the reduction of the wave function. Jammer has previously pointed to
the phenomenological basis of the London}Bauer account (Jammer, 1974, pp.
482}486) but as Gavroglu notes (p. 179), he gets the origins wrong and his
description is thin to say the least. Gavroglu himself gives a detailed analysis
(pp. 169}175) but then concedes too much to Shimony's famous rejection of
what is taken to be an extension of quantum mechanics into psychology
(Shimony, 1963). Shimony's conclusion is that the London}Bauer approach
&rests upon psychological presuppositions which are almost certainly false.'
(ibid., p. 772). However, these presuppositions are not philosophically neutral
and need to be understood from the perspective of London's own philosophy.
Some indication of the latter is given in the Introduction to the monograph
where they write:
Without intending to set up a theory of knowledge, although they were guided by
a rather questionable philosophy [namely the positivistic emphasis on &observable'
quantities], physicists were so to speak trapped in spite of themselves into discovering that the formalism of quantum mechanics already implies a well-de"ned theory
of the relation between the object and the observer, a relation quite di!erent from
that implicit in naive realism, which had seemed, until then, one of the indispensable foundation stones of every natural science (London and Bauer, 1983, p. 220).

Thus in a measurement situation the observer herself must be considered to be


a &system', subject to the theory. What we have then is an ensemble of three
systems, &(object x)#(apparatus y)#(observer z)' (ibid., p. 251) described by the
global wave function
((x, y, z)"& t u (x)v (y)w (z),
(1)
I I I
I
I
where the di!erent states of the observer are represented by the w . For us,
I
considering this combined system as an object, there is little di!erence between
the description of this situation and that in which only the object#apparatus
are considered. The observer, however, has a completely di!erent impression:
For him it is only the object x and the apparatus y that belong to the external
world, to what he calls &objectivity'. By contrast he has with himself relations of
a very special character. He possesses a characteristic and quite familiar faculty
which we can call the &faculty of introspection'. He can keep track from moment to
moment of his own state. By virtue of this &immanent knowledge' he attributes to

 With regard to this section, I am particularly grateful to Matt Taylor for letting me have advance
copies of chapters of his Ph.D. thesis on Husserl (Taylor, 1998).

The Phenomenological Approach to Physics

277

himself the right to create his own objectivity * that is, to cut the chain of
statistical correlations summarized in ((x, y, z)"& t u (x)v (y)w (z) by declar I I
I
I
ing, &I am in the state w ' [2] (ibid., p. 252; London and Bauer's emphasis).
I

The new wave function for the system is therefore not produced by some
mysterious interaction between apparatus and object. Rather, &[i]t is only the
consciousness of an &I' who can separate himself from the former function
((x, y, z) and, by virtue of his observation, set up a new objectivity in attributing
to the object henceforward a new function t(x)"u (x)' (ibid.).
I
According to Shimony two psychological questions must be investigated:
[2] whether mental states satisfy a superposition principle, and whether there is
a mental process of reducing a superposition (1963; p. 760).

He then considers whether a range of psychological phenomena, such as


perceptual vagueness, indecision or con#ict of loyalty, could be interpreted as
instances of superposition, or whether superposition holds in the unconscious;
and he concludes that in both cases the answer is &no'. As for the second
question, the reduction of the wave function could be seen as a result of the
non-causal, creative mental activity of the observer. However, he argues, a) no
more creativity is felt in the case of a quantum measurement than in the case of
a &fully determined' classical one; and b) evolutionary theory makes it di$cult to
understand how irreducibly stochastic behaviour could occur in complex organisms and not in the &primitive entities' at the base of the whole process.
Unfortunately this response mis-characterises the London}Bauer approach
and misses its phenomenological underpinnings. There has, of course, been an
enormous amount of discussion concerning the phenomenological understanding of consciousness in general, and the evolution of Husserl's view of the ego in
particular. However, in the ogical Investigations (Husserl, 1900}1901) (explicitly cited by London and Bauer, together with the Ideas (Husserl, 1913), Husserl
conceives of the ego as just another empirical unity which is constituted by, and
thus only appears in, a re#ective act of consciousness. In the Ideas, and later, in
the Paris lectures, he maintains that this ego can survive the &phenomenological
reduction' but not, according to Taylor (1998) as a Cartesian substance. To say
that the ego appears every time we re#ect on our consciousness is just to say that
there is an ego from the phenomenological perspective (ibid.). The ego can
therefore be understood as &[2] an in"nite cohesion of synthetically connected
acts' (Husserl, 1929, p. 29) and to these acts belong all levels of reality, such as
&ideas' as when we describe nature and the world or treat them theoretically:
In this way we persistently create for ourselves new con"gurations of objects, in
this case ideal objects, which have for us lasting reality. If we engage in radical
self-examination * that is, return to our ego [2] * then all these forms are seen
to be creations of spontaneous &I'-activity [2]. There we also "nd all the sciences,
which, through my own thinking and perceiving, I bring to reality within myself
(ibid., p. 30).
 Gavroglu makes no mention of any further in#uence of Husserl's later work.

278

Studies in History and Philosophy of Modern Physics

The emphasis on our creativity is signi"cant. London and Bauer talk of the
observer attributing to himself &the right to create his own objectivity'. In a typed
addition inserted by London in his own copy of the monograph, he writes:
Accordingly, we will label this creative action as &making objective'. By it the
observer establishes his own framework of objectivity and acquires a new piece of
information about the object in question (London and Bauer, 1983, p. 252).

There is no absolute framework of objectivity residing in some &I' which is


somehow apart from the whole process of observation and which then, by
re#ecting on &its' mental states, collapses the superposition of these states.
Rather the very act of observation itself is a creative construction of objectivity
which &cuts the chain of statistical correlations'. The quantum mechanical
formalism correctly describes the state of the composite object in terms of
a superposition but from &inside' that object, as it were, the observer in the act of
observation creates her own objectivity in the double sense of constructing
the &I' and in doing so, separating this &I' from the composite and thus gaining
the right to choose among the di!erent components of the mixture predicted
by the theory (ibid., p. 251). From this phenomenological perspective the
collapse of the wave function is seen to be part of the two-sided constitution,
through the act of re#ection, of the &I' itself at one pole (Husserl, 1900}1901,
p. 251) and the object at the other. Prior to this act there simply is no &I' to
observe the superposed states of consciousness; it is only in the creative act itself
that both the &I' and object emerge and the collapse occurs.
Now, however, there is an apparent contradiction between this idea of the
pure product state being brought about by an &act of objectifying', and the
apparent ability of measuring arrangements to act as "lters producing pure
cases of the object system (London and Bauer, 1983, p. 257). Reconciliation is
achieved, according to London and Bauer, by noting that although we can
attribute pure states to those particles which have passed through the "lter, we
cannot know which atoms have the property in question since the "lter never
puts any individual object into a new pure state. Thus, they claim, our attribution of the pure state is bought at the cost of the individuality of the object,
which remains absolutely &anonymous' (ibid.). Such a claim is contentious, of
course, but it does mesh with London's earlier insistence, in the work with
Heitler, that electrons, for example, cannot be labelled. London and Bauer
continue by insisting that most measurements are not to do with the properties
of individual systems but rather with general properties of species of systems.
They conclude:
Quantum mechanics, truly a &theory of species', is perfectly adapted to this
experimental task. But given that every measurement contains a macroscopic
process, unique and separate, we can hardly escape asking ourselves to what extent

 As noted above, in fn. 2, London emphasised the &similarity' of electrons in a letter to Born, just
three years before the publication of the monograph with Bauer.

The Phenomenological Approach to Physics

279

and within what limits the everyday concept of an individual object is still
recognizable in quantum mechanics (ibid.).

It is interesting that the term &macroscopic' again crops up here and it plays
a key role in London and Bauer's response to concerns regarding objectivity. At
the beginning of their "nal section, &Scienti"c Community and Objectivity', they
acknowledge that it appears as if quantum mechanics has driven us towards
solipsism. However, they insist, there is still objectivity in the sense of agreement
as to what constitutes the object of investigation. How is this so?
First of all, the act of observation is a &macroscopic', non-quantal act (London
and Bauer, 1983, p. 258). Hence its e!ects on the apparatus can be neglected and
the individuality of the observer can be abstracted away, creating a &collective
scienti"c perception' in which a second observer, looking at the same apparatus,
will make the same observations. However, there is the further worry that the
objects themselves have been reduced to nothing but phantasms produced by
the observer. As they point out, the objectivity of &naive' realism is grounded on
the possibility of continuous connection between the properties of an object and
the object itself, even when it is not being observed. In quantum mechanics that
possibility is no more. Nevertheless we are still able to interpret or predict
experimental results (ibid., p. 259) and this is enough. It is at this point that they
cite Husserl for his systematic study of the necessary and su$cient conditions for
an object of thought to possess objectivity and be an object of science.
This is not to say that the introduction of phenomenology into this context is
unproblematic. Margenau famously criticised it on the grounds that whereas
scientists adopted a fallibilist attitude towards empirical data (and had developed theoretical criteria for the rejection of illusory data), the phenomenologist was guilty of the uncritical admission of introspective evidence which was
regarded as stable and indubitable (and thus had no similar criteria for excluding &abortive introspections' (Margenau, 1950, p. 463). Still, an appreciation of
the phenomenological basis of London and Bauer's approach to the measurement problem is surely necessary for a clear understanding of it.
London and Bauer end their monograph with the remarks that:
One can doubt the possibility of establishing philosophical truths by the methods
of physics, but it is surely not outside the competence of physicists to demonstrate
that certain statements which pretend to have a philosophical validity do not. And
sometimes these &negative' philosophical discoveries by physicists are no less
important, no less revolutionary for philosophy than the discoveries of recognized
philosophers (op. cit., p. 259; London and Bauer's emphasis).

 Husserl insisted that the world is experienced not as our own private world but as an intersubjective one, containing objects accessible to all.
 This is based on his earlier 1944 essay &Phenomenology and Physics', reprinted in Margenau
(1978, pp. 317}328). Margenau was a physics student of London's and went on to work in the theory
of molecular forces (Margenau, 1950, p. xxii; Gavroglu, 1995, pp. 68}69).

280

Studies in History and Philosophy of Modern Physics

These are sentiments which many of us working in the philosophy of physics will
surely agree with.

6. Conclusion
At the beginning of the preface to his book, Gavroglu records that, as he
plunged deeper and deeper into the research, London became a 'very real
person' to him (p. xiii). Rich in detail and insight, Gavroglu's work gives us the
measure of the man along both scienti"c and social dimensions. It stands as
a "tting tribute to a remarkable "gure.
Acknowledgements2Aspects of this review have been discussed with OtaH vio Bueno, Richard
Francks, James Ladyman, Peter Simons and Mauricio SuaH rez. I am grateful to them and to Jeremy
Butter"eld for helpful suggestions. The "nal responsibility for any errors or infelicities rests with me,
of course.

References
Carson, C. (1996) &The Peculiar Notion of Exchange Forces. I: Origins in Quantum
Mechanics, 1926}1928', Studies in History and Philosophy of Modern Physics 27,
23}45.
Cartwright, N., Shomar, T. and SuaH rez, M. (1996) &The Tool Box of Science: Tools for
Building of Models with a Superconductivity Example', in W.E. Herfel et al. (eds),
heories and Models in Scienti,c Processes (Editions Rodopi), pp. 137}149.
French, S. and Ladyman, J. (1997) &Superconductivity and Structures: Revisiting the
London Approach', Studies in History and Philosophy of Modern Physics 28, 263}293.
Heitler, W. (1967) &Quantum Chemistry: the Early Period', International Journal of
Quantum Chemistry 1, 13}36.
Husserl, E. (1900}1901), ogical Investigations, transl. J.N. Findlay (New York: Humanities Press, 1970).
Husserl, E. (1913) Ideas: General Introduction to Pure Phenomenology, transl. W.R. Boyce
(London: Gibson).
Husserl, E. (1929) he Paris ectures, transl. P. Koestenbaum (The Hague: Martinus
Nijho!, 1964).
Jammer, M. (1974) he Philosophy of Quantum Mechanics (New York: Wiley).
Klein, M. J. (1970) Paul Ehrenfest: he Making of a heoretical Physicist (Dordrecht:
North-Holland).
Kuhn, T. S. (1970) he Structure of Scienti,c Revolutions, 2nd edn (Chicago: University of
Chicago Press).
London, F. (1935) &Macroscopical Interpretation of Supraconductivity', Proceedings of
the Royal Society of ondon A152, 24}34.
London, F. (1937) &A New Conception of Supraconductivity', Nature 140, 793}796 and
834}836.
London, F. (1938a) &The j-Phenomenon of Liquid Helium and the Bose}Einstein
Degeneracy', Nature 141, 643}644.
London, F. (1938b) &On the Bose}Einstein Condensation', Physical Review 54, 947}954.
London, F. (1950) Super-uids, Vol. I, (New York: Wiley).
London, F. (1954) Super-uids, Vol. II, (New York: Wiley).

The Phenomenological Approach to Physics

281

London, F. and Bauer, E. (1939) a he& orie de &Observation en Me& canique Quantique
(Paris: Hermann).
London, F. and Bauer, E. (1983) &The Theory of Observation in Quantum Mechanics', in
J. A. Wheeler and W. H. Zurek (eds), Quantum heory and Measurement (Princeton:
Princeton University Press), pp. 217}259.
London, F. and London, H. (1935) &The Electromagnetic Equations of the Supraconductor', Proceedings of the Royal Society of ondon A149, 71}88.
Margenau, H. (1950) he Nature of Phyiscal Reality (New York: McGraw-Hill).
Margenau, H. (1978) &Phenomenology and Physics', in H. Margenau (ed.), Physics and
Philosophy: Selected Essays (Dordrecht: Reidel), pp. 317}330.
Mehra, J. (1994) he Beat of a Di+erent Drum: he ife and Science of Richard Feynmann
(Oxford: Oxford University Press).
Mormann, T. (1991) &Husserl's Philosophy of Science and the Semantic Approach',
Philosophy of Science 58, 61}83.
Ramsey, J. (1997) &Molecular Shape, Reduction, Explanation and Approximate Concepts', Synthese 111, 233}251.
Scerri, E. (1997) &The Case for the Philosophy of Chemistry', Synthese 111, 213}232.
Shimony, A. (1963) &Role of Observer in Quantum Theory', American Journal of Physics
31, 755}777.
Taylor, M. (1998) &Husserl and the Cartesian Epistemological Programme', Ph.D. Thesis,
University of Leeds.

Vous aimerez peut-être aussi