Sie sind auf Seite 1von 40

0039-3681(94)00040-9

Weyl, Reichenbach and the Epistemology of


Geometry

T. A. Ryckman*

‘...physics has in the main contented itself with studying the abridged edition of
the book of nature.’ A. S. Eddington.’

1.
“The Epistemology of Geometry” denotes a Fragestellung which has an
ancient pedigree going back at least to the Pythagoreans. We pick up the story
in a version deriving from Kant, or rather from the demise of Kantian
orthodoxy concerning the synthetic a priori status of geometric concepts, and
extending through Helmholtz and the conventionalisms of Poincare and
Reichenbach. In so far as geometry continues to be held to be “the science of
space”, the problematic of the epistemology of geometry concerns the eviden-
tial basis of the endeavor to determine the geometrical properties of space,
considered as the arena in which physical objects and processes are located. As
is well-known, this problematic is transformed by the use, in general relativity,
of a (pseudo-) Riemannian geometry of variable curvature whose metric in a
given region of space-time is dynamically dependent upon the (local) distribu-
tion of mass and energy. But less well-known are the questions of principle
posed to Einstein’s “solution” to the problem of the epistemology of
geometry, based upon his 1915 theory of general relativity (GTR), by Hermann
Weyl’s attempt to “expand” GTR’s geometrization of gravity into a “unified
theory” encompassing all known forces and interactions. Weyl’s program, in
the period 1918-1924, employed a generalization of Riemannian geometry
expressly designed to challenge the congruence assumptions underlying
Riemannian geometry (and hence GTR). Significantly, Weyl’s criticisms arose
from an epistemological perspective according to which these assumptions
*Department of Philosophy, Northwestern University, Evanston, IL 60208-1315, U.S.A
Received 26 March 1993; in final form18 March 1994.

“A Generalization of Weyl’s Theory of the Electromagnetic and Gravitational Fields’, Proceed-


ings of the Royal Society of London A99 (1921), lOk-127; see p. 108.

Pergamon Stud. Hist. Phil. Sci., Vol. 25, No. 6, pp. 831-870, 1994
Copyright 0 1995 Elsevier Science Ltd
Printed in Great Britain. All rights reserved
0039-3681/94 $7.00+00.00

831
832 Studies in History and Philosophy of Science

appeared as an entirely avoidable blemish in Riemann’s own naturphiloso-


phische aspirations; in the context of GTR, Weyl’s views fundamentally
question the legitimacy of Einstein’s tacit assumption of the existence of
“practically rigid bodies” to which the central theoretical structure of GTR,
the metric interval ds, was “normed”, i.e. held to correspond.
In brief, Einstein had, as a practical expedient, followed Helmholtz in
postulating the existence of de facto rigid bodies corresponding to metrical
concepts (in particular, congruence).2 To be sure, Einstein was well aware that
acute thinkers, notably Poincare, had demonstrated that the use of physical
objects and processes as correlates for geometrical concepts was not at all
innocent, and that one could thus only arbitrarily attribute geometrical
properties to space itself as distinct from the space implicated by the behavior
of these objects and processes. Poincare’s assessment, Einstein would note (see
Section 4 below), was unimpeachable sub specie aeterni. Indeed, this lofty
vantage point, momentarily occupied already by Weyl, is none other than that
of a “systematic theory”, a unified theory of fields, to whose discovery Einstein
would unsuccessfully devote the largest portion of the remainder of his life. Yet,
at least for the moment, for Einstein the empirical basis of GTR rested upon the
coordination of the ds to “infinitesimal rigid bodies” or material processes
(atomic spectra), a supposition which, if not completely theoretically satisfying,
was in perfect accord with the observable facts about transported measurement
bodies.
In any case, this explicit acknowledgement of the seemingly unavoidable role
of physical indicators (“rigid bodies”, “test particles”, “atomic clocks”, light
rays) in providing an evidential basis for belief in a particular geometrical
characterization of space, pointed to an apparent need for a clear distinction
between “pure” and “applied” geometries, and for an epistemological account
for the former’s relation to the latter. By the first decade or so after the
inception of general relativity in 1915, there had emerged two scientifically
respectable responses to these new developments in the problem of the
epistemology of geometry.
The first, building upon the climate of pluralism created by the 19th century
discovery of new taxa of geometry, notably, projective and non-Euclidean
geometries, developed a purely formal axiomatic conception of geometry.
Associated primarily (and somewhat unfairly) with the name of Hilbert,’ it

‘The affiliation with Helmholtz is established by Einstein himself in a little-known essay,


‘Nichteuklidische Geometrie und Physik’. LXe Neue Rundschau 36 (1925), 1620. I’d like to thank
Don Howard for calling my attention to this essay. Einstein’s (and Reichenbach’s, see Section 4
below) association with the geometric views of Helmholtz is surprising, given the fact that
Helmholtz’s conception permits only geometries of constant curvature, not the geometries
(Riemannian and the Weyl-Eddington non-Riemannian varieties) of variable curvature of GTR.
?See Hans Freudenthal, ‘The Main Trends in the Foundations of Geometry in the 19th Century’,
in E. Nagel, P. Suppes, and A. Tdrski (eds), Logic, Methodology and Philosophy qf Science;
Proceedings of the 1960 Confrrence (Stanford, CA, Stanford University Press, 1962) pp. 613321.
Weyl, Reichenbach and the Epistemology of Geometry 833

becomes part of the epistemological orthodoxy of logical empiricist philosophy


of science due to Einstein’s famous maxim in his lecture “Geometry and
Experience” to the Prussian Academy in January 1921:
. ..as far as the propositions of mathematics refer to reality, they are not certain; and
as far as they are certain, they do not refer to reality.4

Geometry, under this purely formal axiomatic conception, becomes simply “a


mathematical science”, i.e. “a body of theorems deduced from a set of axioms”.
As to why the name “geometry” is given to some mathematical sciences and not
to others, Oswald Veblen, an important proponent of this view, provided the
only reasonable answer:
(B)ecause the name seems good, on emotional and traditional grounds, to a sufficient
number of competent people.5

The governing conception here is a complete rejection of the “traditional”


viewpoint, according to which geometry singles out, by privilege of intuition,
any particular subject matter, notably, that of space; that this has been true of
geometry in the past is simply an artifact of culture.
Since geometry has thus been eviscerated of all non-formal content, we arrive
rather naturally at the epistemological juncture intended by Reichenbach to
canonically summarize Einstein’s views on the relation of mathematical
concepts to the physical world: the purely formal objects of mathematics
(geometry) are “coordinated to” (zugeordnet) physical objects6 the paradigm
case being the coordination of metric concepts to measuring rods and clocks.
As could be argued at length,’ in adopting the language of “coordination” to
characterize the relation of mathematical concepts to physical entities in
theoretical physics, both Reichenbach and Einstein show the influence of the
formalist epistemology of Schlick’s Allgemeine Erkenntnislehre of 1918. Here,
while conscientiously rooting out any cognitive role for, or reliance on, notions
of intuition, Schlick had proposed an account of scientific cognition wholly in
4’Geometrie and Erfahrung’, Sitzungsberichte. Preussische Akademie der Wissenschaften,
Physikalische-Mathematische Klasse, 1921, pp. 123-130; separately issued in expanded form and
translated by W. Perret and G. B. Jeffrey, as ‘Geometry and Experience’, in Sidelights on Relativity
(New York, E. P. Dutton, 1923) pp. 27-56; see p. 28.
s0swald Veblen and J. H. C. Whitehead, The Foundations of Differential Geometry (Cambridge,
Cambridge University Press, 1932) p. 17.
6‘Geometrie and Erfahrung’ (op.cit., note 4). p. 125: “It is clear that the system of concepts of
axiomatic geometry alone cannot make any assertions as to the relations of real objects...which we
will call ‘practically-rigid’ bodies. To be able to make such assertions, geometry must be stripped
of its merely logical-formal character in that empirical (erlebbare) objects of reality must be
coordinated (zugeordnet) to the empty conceptual schema of the axiomatic geometry.”
‘See Don Howard, ‘Einstein, Kant, and the Origins of Logical Empiricism’, in W. Salmon and
G. Wolters (eds), Language, Logic, and the Structure of Scientt$c Theories (Pittsburgh, University
of Pittsburgh; Konstanz, Universitat Konstanz, forthcoming), and, on Schlick’s epistemological
employment of this concept, my ‘Conditio Sine Qua Non? Zuordnung in the Early Epistemologies
of Cassirer and Schlick’, Synthese 88 (199 I), 57-95.
834 Studies in History and Philosophy of Science

terms of implicitly defined concepts and of their Zuordmmg to real objects or


objects of experience. The Hilbertian resonances of his characterization of
cognition are pointed out by Schlick himself, who also claimed to have thus
provided a complete refutation of Kantian epistemology. We might note in
passing that precisely this sharp distinction between formal science and
empirical content was to become the hallmark of logical empiricism under the
rubric of a sharp analytic/synthetic distinction, a distinction in which, unlike in
Kant, there is no longer any cognitive role for intuition, either pure or
empirical.
The other approach differs perhaps more in ideology or epistemological
orientation than in other respects. Its leading figures are Riemann, Lie, Klein,
Weyl and Cartan; its main mathematical tenet is not the formal axiomatic
approach to geometry but rather its continuous (i.e. Lie) group-theoretic
characterization.* While not committing the original Kantian sin of positing a
necessarily Euclidean structure to the space of intuition as a condition of
experience, neither does it abjure talk of intuition and space altogether. Indeed,
the preference for local (i.e. differential and tangent space) characterizations
governed the approach here to what was still referred to as “the problem of
space” (Das Raumproblem), although, at least in Weyl’s terms, this is now to
be conceived as the problem of space itself, not merely a doctrine of the
configurations possible in space. 9 Naturally, such an approach is not math-
ematically incompatible with a purely axiomatic approach, but epistemologi-
tally, it is quite at variance with the axiomatic viewpoint in its insistently local
chauvinism. Here a kind of naturphilosophische demand is made that the world
be understood in terms of its behavior in the infinitely small; only there will be
found the simple elementary laws which genuine understanding requires.‘O In
Weyl’s case, the epistemological advantage of such an approach is obvious:
only the spatially-temporally coincident and the immediate surrounding

‘On the classification, see 8. Cartan, ‘Le R6le de la Thborir des Groupes de Lie dans L’bvolution
de la GkomPtrie Modern’, Comptes Rendus de la CongrPs internationale, Oslo 1936 I, pp. 92.-103;
reprinted in Oeuvres Complgtes, 2nd edn, partie III 2 (1984), 1373-1384. It is Cartan who will
employ Lie groups most broadly as a tool of unification in geometry with his theory of “generalized
spaces”: roughly, a space of tangent spaces (a.k.a. a “connection”) such that two infinitely near
tangent spaces are related by an infinitesimal transformation of a given Lie group.
‘Space-Time Matter (op.cit., note 26 below), p. 102.
‘“For example, ‘The Lie theory [of continuous groups] belongs in that great train of thought
(Gedankenzug) which would understand the world from its behavior in the infinitely small and
which has proved itself so fruitful because only the passage back to the infinitely small leads to
simple elementary laws, Its standpoint is intuitively thoroughly natural; for example, instead of
describing the mobility in Euclidean space of a rigid body rotating about a fixed point according
to the requirement that the location of the body at each moment proceed from its initial location
through an operation of the Euclidean rotation group, it is natural to conceive of the body’s
continuous rotation as an integral succession (Aneinanderreihung) of infinitesimal operations of this
group, which follow upon one another according to the single elements of time.’ Hermann Weyl,
‘Kontinuierliche Gruppen und ihre Darstellungen durch Lineare Transformationen’, in Atti de1
Congress0 Internazionale dei Marematici, Bologna 3-10 Settembre 1928 I, pp. 233-246; see p. 243.
Weyl, Reichenbach and the Epistemology of Geometry 835

neighborhood at once have a clearly exhibited meaning in intuition. *l Intuition


(Anschauung) here has a linkage to “immediate insight” or “evidence” that
reflects, loosely speaking, rather more of an Husserlianiz than a Kantian
nuance. Indeed for Weyl, since Evidenz “is the sole source of all insight”, the
formalist epistemology of Schlick will simply remain “not understandable”.‘3
Weyl’s epistemological objections to formalism are further revealed, of course,
in the foundations of mathematics. In the very same year (1918) in which he
publishes the initial papers on his unified theory of gravitation and electromag-
netism (see below), Weyl will also seek to counter the trend toward formalism
associated with his teacher Hilbert, basing even the justification of mathemati-
cal propositions not on proof, but on “immediate insight” which furnishes
“the experience of truth”.14 But no less a striking contrast, elucidated below,
shows up in the foundations of GTR, a distinction that Einstein himself
encapsulated in another context in 1919 in pointing to the difference between
kinds of theories in physics: thus the general theory of relativity, is, on the one
hand with Pauli, Reichenbach and Einstein (at least in this period of his career),
viewed as a theory ofprinciple whose basis lies in empirical fact, or on the other,

“‘Nur das raumzeitliche Zusammenfallen und die unmittelbare raumzeitliche Nachbarschaft


haben einen in der Anschauung ohnes weiteres Klar aufweisbaren Sinn.’ ‘Geometrie und Physik’
(The Rouse Ball Lecture at Cambridge University, May, 1930) Die Nuturwissenschaften 19 (1931),
49-58; see p. 49.
“I cannot here give Weyl’s tortuously complex epistemological views the attention they deserve,
but two points may be insisted upon: (1) to label Weyl an “Husserlian” (as is often done) does little
to further understanding; (2) Weyl’s own research activity in the exact sciences provides the arena
for interpretation of his epistemological reflections. Consider the relationship between these two
passages: “The world comes only into consciousness in the general form of the consciousness which
is there (welche da ist); a penetration of Being and essence (Durchdringung des Seins und Wesens),
of the ‘this’ and ‘thus’ (‘Dies’ und ‘So’). (The intimate understanding of this penetration, it may be
remarked, is according to my conviction the key to all philosophy)” [“Das Verhaltnis der kausalen
zur statischen Betrachtungsweise in der Physik’, Schweizerische Medizinsche Wochenschrzft 34
(1920), 131-741; see p.738.1 ‘In$eld theory the space-time continuum plays in a certain sense the role
of substance, if we conceive of the opposition of Substance and Form as that of the ‘this’ and ‘thus’
(‘Dies’ und ‘So’). . ..According to field theory the world description consists, to use the terminology
of Hilbert, in “here - so relations” - the ‘here’ represented through the space-time coordinates, the
‘so’ through the state magnitudes.’ Was ist Materie? Zwei Aufsiitze zur Naturphilosophie (Berlin,
Springer, 1924) p. 42.
‘%ee Weyl’s review of Schlick’s ANgemeine Erkenntnislehre (1918) in Jahrbuch iiber die
Fortschritte der Mathematik 46 (1923), 59-62: esp. p. 60: ‘To the reviewer it is not understandable
how anyone who has ever attained an insight (Einsicht) can be satisfied with this point of view [i.e.,
with Schlick’s conception of the essence of knowledge as lying in the merely designative or
“semiotic” conception of concepts-TR].... In as much as [Schlick] ignores intuition, in so far as it
goes beyond the merely perceptual-like sensuous, in such wide measure, he frankly discards
Tverwirf;) evidence (Evi>eni) which is still the sole source of all insight.’
14Das Kontinuum (Leiuzig, Veit. 1918). n. 11. footnote: “The famous Dedekind essay Was sind und
was sollen die Zahlen begins (Foreword to the 1st edition) with the statement: ‘What is proveable
should not be believed in science without proof’. This remark is certainly characteristic of the mode
of thinking of most mathematicians; nevertheless, it is a preposterous principle. As if such an
indirect collection of grounds, though we call it a ‘proof’, is capable of awakening any ‘belief
without our assuring ourselves in immediate insight (unmittelbarer Einsicht) of every single step.
This (and not the proof) generally remains the last source of justification of knowledge; it is the
“experience of truth” (Erlebnis der Wahrheit).”
836 Studies in History and Philosophy of Science

with Weyl, as a constructive theory, built up from a hypothetical and ideally


simple starting point.i5
In this paper, I will seek to elucidate this difference by contrasting the views
of Weyl and Reichenbach on what can thus, only with some ambiguity, be
called “the epistemology of geometry”. To do so, I will situate the develop-
ment of Reichenbach’s geometric conventionalism as occurring against the
backdrop of, and in no small measure in response to, an explanatory strategy
adopted by Weyl to defend his broadening of the general theory of relativity,
incorporating the electromagnetic field into the metric of space-time, against
what appeared at the time to be a decisive objection raised by Einstein. In
particular, my aim is to show that the central mechanism by which Reichenbach
created his metaphilosophical theory of “equivalent descriptions” (destined to
become part of the methodological edifice of logical empiricist philosophy
of science) is, in its appeal to a little-understood distinction between
“differential” and “universal” forces, targeted on Weyl’s theory. More exactly,
it is directed against Weyl’s explanatory account of how it is that we
do observe the congruence behavior of transported measuring rods (or the
constancy of frequencies of spectral lines of atoms) that we do, despite the
existence of the central theoretical structure in this unified theory, a space-time
with a non-Riemannian metric, according to which vector transference is not
generally integrable. But beyond establishing such an historical linkage, this
case study aims to demonstrate how an influential and exceedingly general
(“global”) characterization of theoretical underdetermination arose as a one-
sided, hence questionable, extrapolation from a critical foundational contro-
versy within the nexus of the revolutionary ideas of GTR itself: a story that
should serve to renew our skepticism of this manner of practising philosophy of
science.
In what follows, a summary exposition of Weyl’s theory stressing its
philosophical motivations precedes (in Section 3) an account of Weyl’s reply
to the initial criticisms of Einstein and Pauli that his theory conflicts with
observation. We then turn in Section 4 to an examination of the somewhat
tempered position of Einstein in his lecture of January, 1921, “Geometry
and Experience”, contrasting this with the response of Reichenbach as
tracked through the emergence of his characteristic form of geometric
conventionalism.

“The distinction is drawn by Einstein in an essay ‘What is the Theory of Relativity?’ at the
request of The Times of London, 28 November, 1919, reprinted in Ideas and Opinions (New York,
Bonanza Books, 1954) pp. 227-232. I have been prompted to apply the distinction here by the
discussions (which point to Einstein’s switch, in the case of GTR, from the former to the latter) in
two papers of John Stachel, ‘Special Relativity from Measuring Rods’, in R. S. Cohen and
L. Laudan (eds) Physics, Philosophy and Psychoanalysis; Essays in Honor of Adolf Griinbaum,
(Dordrecht, D. Reidel, 1983) pp. 255-273; and ‘Einstein and the Quantum: Fifty Years of
Struggle’, in R. Colodny (ed.), From Quarks to Quasars: Philosophical Problems of ModernPhysics,
(Pittsburgh, University of Pittsburgh Press, 1986) pp. 3499385.
Weyl, Reichenbach and the Epistemology of Geometry 837

Hermann Weyl’s “expansion” of the theory of general relativity in 19 18-l 923


was, following the inception of general relativity by Einstein in 1915-1916, the
first and historically most important attempt to offer on its basis a “unified
theory” of all known physical forces, at that time including only gravitation
and the forces of the electromagnetic field. 16 In so doing, Weyl inaugurated
efforts which were to occupy Einstein, unsuccessfully, until his death, and
created a tool, “local gauge invariance” which, in somewhat different form,
continues to be employed by contemporary physics of elementary particle
interactions. What is not widely remembered today is that Weyl, who was from
19 16 one of the most prominently strong supporters of GTR, had a philosophi-
cal understanding of GTR that was not only distinctively different, in crucial
places, from that of Einstein,17 but also completely at variance with the received
philosophical interpretation of GTR by logical empiricism which was largely
the creation of Hans Reichenbach and which, until recently, was passed down
to subsequent generations of philosophers of science as a lasting achievement,
part of the canon of scientific philosophy.18
Weyl’s unified theory, which Weyl himself actively promulgated until
the mid-1920s and the advent of Quantum Mechanics, is really only
16Hilbert’s famous “notes” on the foundations of physics of November 1915 and December 1916
(Kiinigl. Gesellsrhafi d Wissenschaften zu Giittingen. Nachrichten. Math.-Phys. Klasse 1915,
pp. 396407; 1917, pp. 53-76) are not “unification” attempts in the same sense, since although
Hilbert (using invariant theory and a variational method) derived Mie’s equations of electrody-
namics from the field equations of gravitation (independently discovered by Hilbert in November,
1915) the Mie theory is valid only in the limiting case of special relativity; for discussion, see
J. Mehra, Einstein, Hilbert, and the Theory of Gravitation (Dordrecht/Boston, D. Reidel, 1974)
pp. 2430.
“A very fundamental difference concerns the interpretation of the relativity of motion where
Weyl takes issue with Einstein’s “relativist” response to Lenard at the Bad Nauheim Naturforscher
Versammlung in September 1920: ‘If, with Einstein, one can adapt a coordinate system to all
particles such that simultaneously one can transform them all to a state of rest, it no longer makes
sense to speak of relative motion. . ..The principle that matter generates the field can therefore only
be straightforwardly maintained if the concept of motion admits in itself a dynamic moment. The
analysis of the concept of motion revolves not around the opposition absolute or relative but rather
around kinematic 0; dynamic.’ Die Relativitiitstheorie auf her Naturforscherversammlung in Bad
Nauheim, Jahresberichte der Deutschen Mathematische- Vereiniaunp (1922). DD. 51-63 and 62-63;
also, ‘As long as one ignores the guiding field (Fiihrungsfeld), one can speak neither of relative nor
of absolute motion; only with regard to the guiding field does the concept of motion gain a content.
Relativity theory, correctly understood, will not eliminate absolute motion in favor of relative, but
it denies (vernichtet) the kinematic conception of motion and replaces it through the dynamic.’
‘Massentragheit und Kosmos,’ in Zwei Aufsiitze zur Naturphilosophie (op.cit., note 12) p. 66; on the
concept of “guiding field”, see Section 2 below.
“See especially, J. A. Coffa, ‘Elective Affinities: Weyl and Reichenbach’, in W. Salmon (ed.),
Hans Reichenbach: Logical Empiricist (Dordrecht, D. Reidel, 1979). pp. 2677304. I would like
to record here my debt for the stimulation provided by this paper, the only one in the literature on
its subject matter. Coffa focuses on the conflict of the respective viewpoints regarding the
interpretation of relativity (in particular, relative motion), expressly leaving aside “their disagree-
ment over the empirical determination of the metric”, which is our topic. Coffa modestly said his
paper is “no more than an invitation to pursue a topic” (p. 267) but it is certainly much more than
that.
838 Studies in History and Philosophy of Science

understandable as a bold, but ultimately unsuccessful attempt to remove, from


a perspective that could be described as Riemann’s own, a Schoenheitsfehler’9
from Riemann’s theory of manifolds, the mathematical framework in whose
terms GTR is cast. This framework, it should be noted, uneasily coexists within
the problematic of “the epistemology of geometry”: for indeed it transforms the
very meaning of “geometry” as the term applies to the theory of physical space.
And it is this transformed conception of geometry, in the current consensus,
that has won the day in GTR.
Weyl’s theory, or at least its motivation, really begins with Riemann’s famous
1854 Hubilitationsschrif “On the Hypotheses which Lie at the Foundations of
Geometry”.20 Here Riemann sketched, in briefest outline, an approach to the
mathematical investigation of the concept of space by conceiving of spatial
objects, which he called “multiply extended magnitudes” or, what we know
today as n-dimensional topological manifolds. For Riemann, physical space,
the space of our encounters with external objects, must be initially considered
as merely a special case of a triply extended (3-dimensional) manifold. The
concept of extension in itself determines no particular geometry or system of
metric relations. Of course, Riemann did not ignore the question of metric
relations in space. He allowed that various systems of geometric axioms were
possible but that none was logically (or for that matter, cognitively) necessary,
and each was empirically contingent in its description of space. Tantalizingly,
he even questioned the continuity of space, suggesting, in an obscure passage,
that if space is in fact continuous, its metric relations may stem from something
external to space, namely, matter. And he speculated that the viability of
familiar metric concepts, such as the rigidity of measuring rods, may break
down in the domain of the infinitely small. 21 The mathematical significance of
Riemann’s address is, of course, of the highest order, containing as it does in
embryo, the foundations of differential topology and geometry. But it is not an
overestimation of Riemann’s paper, scarcely 20 pages long with very few

“This characterization is Dirk J. Struik’s, ‘Schouten, Levi-Civita, and the Emergence of Tensor
Calculus’, in David E. Rowe and John McCleary (eds), The History of Modern Mathematics,
Vol. 2 (Boston and New York, Academic Press, 1989). DU. 98-108: see D. 104.
“Largely unknown to English readers, Weyl also”e&ed an edition of Riemann’s essay, B.
Riemann, Ueber die Hypothesen, welche der Geometrie zu Grunde liegen, Neu herausgegeben und
erltiutert van H. WeyI (Berlin, J. Springer, 1919); Dritte Auflage, 1923. Weyl’s editorial
“elucidations” of Riemann’s text are an important source for his own epistemological and
noturphilosophische views,
“‘But now it seems that the empirical concepts, in which spatial measure-determinations are
grounded, the concepts of rigid body and light ray, appear to lose their validity in the (domain of
the) infinitely small. It is therefore easily imaginable that the measure-relations of space in the
infinitely small are not in accord with the presuppositions of geometry and in fact this must be
supposed as soon as the phenomena might thereby be more simply explained.’ Concerning this
passage (at pp. 19-20), Weyl observed: ‘The complete understanding of the concluding remarks of
Riemann concerning the underlying basis of the metric relations of space (inner Grund der
Massverhiiltnisse des Raumes) was first disclosed to us by Einstein’s general theory of relativity’,
ibid., pp. 4641.
Weyl, Reichenbach and the Epistemology of Geometry 839

formulas, to attribute to it an equal philosophical importance.22 For, by


recognizable intention and with the advantage of hindsight, it also provides the
resources for a complete refutation of Kant’s doctrine of space. But this can
perhaps best be seen in the work of Weyl on general relativity we are about to
consider.
In his own mathematical development of these ideas, written a few years
later, in 1861, for a prize competition in Paris (he didn’t win!), Riemann
more or less conjured up an expression (a quadratic differential form) which
is a generalization of the familiar Pythagorean theorem and which gives
metrical relations in such a manifold.23 It contains, in addition to the
differentials of the coordinates, functions of the coordinates as coefficients to
the coordinate differentials which in the general case may have different
values at different points of space. Such an expression provides a metric not
only for curved spaces, but also for spaces of variable curvature. Riemann’s
ideas about variably curved space were indeed revolutionary; no less a
mathematician than Henri Poincare, writing in the early 1890s and again in
his well-known Science and Hypothesis of 1903, expressed the judgement that
these spaces were but mathematical curiosities and, in his words, “could
never be other than purely analytic”, that is, not axiomatizable, hence
unsuitable candidates for geometry in any conventionally chosen combi-
nation of geometry and physics as a theory of physical space.24 In his 1861
prize essay, Riemann had considered only the case of manifolds in which the
direction, but not the length or magnitude of a vector, was altered under

**The assessment here relies on the case made in Gregory Nowak, ‘Riemann’s Habilitarionsvor-
rrag and the Synthetic A priori Status of Geometry’, in David E. Rowe and John McCleary (eds),
The History of Modern Mathematics, Vol. 1 (Boston and New York Academic Press. 1989).
pp. 1748.
“See R. Farwell and C. Knee, ‘The Missing Link: Riemann’s Commentario, Differential
Geometry and Tensor Analysis’, Historia Mathematics 17 (1990), 223-255. A standard history
writes that the Riemann expression of the distance element d.?=C,,g,,dx’dxk was, for Riemann,
“an article of faith, essentially”; J. A. Coolidge, A History of Geometrical Methods (Oxford,
Clarendon Press, 1940), p. 410. Of course, its validity represents the supposition of the validity
of the Pythagorean theorem in the domain of the “infinitely small”; Weyl will provide a proof of
the “uniqueness” of the Pythagorean measure of the line element in this domain, which
establishes that it is the only measure under which the (infinitesimal) orthoeonal
- groun is
-I
a volume-preserving invariant, ‘Die Einzigaragkeit der Pythagoreischen Massbestimmung’,
Mathematische Zeitschrift 12 (1923), 114146; a slightlv more readable oresentation is in
Mathematische Analyse des Raumproblems; Vorlesun~en- gehalten in Barcelona und Madrid
(Berlin, J. Springer, 1923) lectures 7 and 8; see the brief discussion towards the end of Section 2
below.
Z4‘Les Geometries non-euclidiennes’, Revue g&&ale des Sciences pures et appliqukes 2 (1891)
7699774; this essay is reprinted as chapitre III of La Science et L’HypothPse (Paris, 1903;
reprinted by Flammarion, Paris, 1968), cited passage at p. 73: ‘Ces geometries de Riemann, si
interessantes a divers titres, ne pourraient done jamais &tre que purement analytiques et ne se
preteraient pas a des demonstrations analogues a celles d’Euclide’. Incidentally, the only in-print
English translation of this work (by W. J. Greenstreet, published by Dover, New York) garbles
this passage (at p. 48) completely inverting its meaning, as was already pointed out by Bertrand
Russell in his review of this translation, Mind 14 (1905), 412418.
840 Studies in History and Philosophy of Science

transport in the manifold.25 In the general case, transporting a vector around


a closed curve and returning it to its point of origin alters its direction but
not length; hence, length is integrable, but direction is not. Such manifolds,
possessing as metric the positive-definite quadratic form mentioned above,
are known as Riemannian manifolds.
All very well, but what has this to do with Weyl? As Felix Klein had done
some twenty years before, Weyl recognized in Riemann’s differential geometry,
which Weyl was soon to generalize, a fundamental epistemological postulate:
that genuine understanding or explanation of the processes of nature studied in
physics is produced only through conceptions regarding the behavior of nature
in the infinitely sma11.26 For Weyl, Riemann’s passage from a Euclidean
“distant” (“Fern”) to a Riemannian “local” (“N&e”) geometry is exactly
parallel to the transition in physics from action-at-a-distance theories to the
physics of fields. We can understand, e.g. Coulomb’s law or Ohm’s law which,
in Weyl’s terms, are Fernwirkunggesetzen (“action-at-a-distance laws”), only
because these are derivable from Maxwell’s equations of the electromagnetic
field. Similarly, Euclidean geometry formulated to meet the requirements of
continuity is Riemannian geometry, in which there is a “local” Euclidean
structure in the neighborhood of each point. The transition from Euclidean
“distant” geometry to Riemannian differential geometry (and, as we shall see,
on to a non-Riemannian “pure infinitesimal geometry”), as well as the passage
from action-at-a-distance physics to a physics of action by immediate contact
(Nuhewirkungsphysik) both illustrate the epistemological gain in understanding
brought by restricting one’s considerations to behavior in the infinitely small.
Why is this? Because in the passage to the infinitely small, all problems can be
linearized: this is the approach of the differential calculus, infinitesimal
25That Riemann considered only this restricted case, corresponding to, in Weyl’s theory, the
absence of an electromagnetic field (F,k=O, see below) was due to the historical development of
Riemann’s views in Gauss’s theory of surfaces; see Weyl’s letter to Einstein of 19 May 1918 (cited
in N. Straumann, Zunz Ursprung der Eichtheorien bei Hermann WeyL Physikalische Bliitter 43
(1987), 414-421, at p. 416. For an extended discussion of this development, see Weyl’s edition of
Riemann’s essay (op.&. , note 20), pp. 3538.
26’The principle, to understand the world from its behavior in the infinitely small, is the driving
epistemological motive of contiguous-action physics (Nahewirkung physik) as well as Riemannian
geometry, but it is also the driving motive in the grandiosely directed remaining life-work of
Riemann, above all, the theory of complex functions.’ Raum-Zeit&Materie (Berlin, J. Springer,
1918) p.82, repeated through the 5th edition (1923) p. 86, of this work; the 4th edition, translated
into English as Space-Time Matter by Henry L. Brose (London, Meuthen and Co. 1923; reprinted
by Dover, 1952) translates the passage somewhat differently on p. 92. In a Vienna lecture in 1894,
Klein had made the same observation, linking the names of Riemann and Faraday: ‘As in physics,
the banishment of actions-at-a-distance has led to the explanation of the phenomena through the
inner forces of a space-filling aether, so in mathematics, the understanding of functions arises from
their behavior in the infinitely-small, especially, therefore from the differential equations which
satisfy them. . ..If I might dare to so sharply press the analogy, then I would say that Riemann in
the domain of mathematics and Faraday in the domain of physics stand parallel.’ ‘Riemann und
seine Bedeutung fur die Entwicklung der modernen Mathematik’, Jahresbericht der
Deufschen Mathematiker- Vereinigung, Vol. 4 (18945) equivalent to Gesammelte Mathematische
Ahhandlungen, Vol. 3 (Berlin, J. Springer, 1923) pp. 482497, at p. 484.
Weyl, Reichenbach and the Epistemology of Geometry 841

geometry and field physics.27 Linearizing promotes mathematical tractability,


but thereby we also gain in intuitive clarity, we come to understand.
But there is also an interesting and unusual connection between the
epistemological requirement to understand nature through the infinitely small
and Weyl’s version of holism, i.e. to the inseparability of physics and geometry
and to the impossibility of verifying single statements independently of the
nexus of theoretical commitments in which they are embedded. In an example
he repeatedly uses, Weyl argues that it is field physics (or rather the Wechsel-
wirkung between matter and field) which illustrates the unavoidability of
epistemological holism in fundamental physics, that is the physics of the
interactions of matter and field. The example concerns, on the one hand, the
so-called “ponderomotive” force exerted upon a test charge by the electro-
magnetic field, and on the other, the generation of field strengths by the test
charge. The laws and interactions here constitute a cycZe: the field acting upon
the test charge, which in turn disturbs the surrounding field. Weyl’s use of this
example occurs in the context of arguing that physics and geometry are an
inseparable whole.28 Dynamical Wechselwirkung (interaction) is the character-
istic mode of action transmission in field theories. Einstein had of course shown
that the inertial and gravitational structure of the world is not geometrically
rigid but active and dynamic; it is an interactive field like the electromagnetic
field and this analogy always governs Weyl’s presentation of general relativity.
Just as the Wechselwirkung between a charged body and the surrounding
electrical field can only be understood in terms of the “contiguous-action”
(Nahewirkung) represented in the partial differential equations of Maxwell’s
theory, so also the reciprocal influence between what Weyl was to term the
Fiihrungsfeld (“guiding field”, the inertial-gravitational structure of the world)
and a massive body within it calls out for explanation in the same terms. As we
shall see, from a “purely infinitesimal standpoint” of differential geometry and
the “contiguous-action” of field physics, the apparently congruent behavior of
transported measuring rods should not be accepted as brute fact-to do so
shows a complacency toward the remnants of the old Euclideanferngeometrisch
analysis of nature. Weyl will thus demand that the behavior of measuring rods

27Mathematische Analyse des Raumprohlems (op.cit., note 23), pp. 17-18.


*‘Raum-Zeit-Materie, 1918 (op.cit., note 26) p. 60 (and repeated verbatim in all subsequent
editions): Space-Time+Matter, p. 67: ‘The distribution of the elementary quanta of matter provided
with charges fixed once and for all (...) determine the field. The field exerts upon charged matter a
ponderomotive force. . ..The force determines, in accordance with the fundamental law of
mechanics, the acceleration, and hence the distribution and velocity of the matter at the following
moment. We require this whole network of theoretical considerations to arrive at an experimental test
~ if we assume what we directly observe is the motion of matter. . ..We cannot merely test a single
law detached from this theoretical fabric! The connection between direct experience and the
objective element behind it, which reason seeks to grasp conceptually in a theory, is not so simple
that every single statement of the theory possesses an immediate, intuitively verifiable meaning. We
will . ..see that geometry, mechanics and physics comprise in this manner an inseparable theoretical
unity...‘.
842 Studies in History and Philosophy of Science

and clocks be derived from a “systematic theory”, i.e. a fully dynamical theory
of matter-field interactions.
Weyl was among the first strong supporters of the general theory of
relativity,29 and one of the first to lecture on general relativity, which he did
in the summer of 1917 in Zurich where he was professor of mathematics.
Weyl was then 31 years old and had just been released from the German
army into which he had been drafted in 1915. Upon returning to Zurich, he
had immediately immersed himself in Einstein’s theory while looking for
something to whet his mathematical curiosity.30 These lectures were written
up and published in the spring of 1918 under the title Raum-Zeit-Materie;
the book was to go through five editions by 1923. The fourth edition was
(rather poorly) translated into English in 1922 and is still in print, while a
German edition (the 7th) is also still in print. This was the first textbook (if
it can be called that!) on general relativity and it won high praise from
Einstein, who wrote in a rare book review that “every page reveals the
unerring hand of the master”.” In the first edition, Weyl had not yet
introduced his “expanded” theory and he largely adhered to the geometrical
development of general relativity which followed Einstein’s own exposition in
proceeding from the basic structure of a four-dimensional Riemannian
manifold with a special case of the Riemann metric, called a Lorentz metric,
which takes account of the light cone (hence, causal) structure of space-
time.32 But no sooner was the book out then Weyl, spurred on by a 1917
paper by the Italian geometer, Tullio Levi-Civita, conceived of a stunning
generalization of Einstein’s theory. As great as his respect was for Einstein’s
theory, Weyl now believed, from the naturphilosophische standpoint of the

29Writing to Weyl on 23 November 1916, Einstein remarked, clearly with Weyl (and Hilbert) in
mind, that ‘the average mental power of its [i.e., GTR’s] supporters by far surpasses that of its
opponents. This is a kind of objective evidence for the naturalness and sensibleness of the theory’,
translated and quoted in Sigurdsson (op. cit., note 30 below), p. 161.
30Biographical details can be found in Skuli Sigurdsson, Hermann Weyl, Mathematics and
Physics, 1900&1927, Harvard Ph.D. dissertation, 1991.
3’Deutsche Literaturzeitung, Vol. 25 (21 June, 1918), Columns 372-373. Einstein’s review of the
first edition is remarkable for its effusive praise; he commends Weyl in particular for his success in
integrating the field equations of gravitation, a treatment showing “how simplifying and clarifying
is the work of the born mathematician there”. Interestingly, towards the end, Einstein expresses
some reservation concerning the conception Weyl gives of the relations obtaining between the
expressions of theoretical physics and reality: ‘ich mit dem Verfasser nicht ganz iibereinstimme
beziigich der Auffassung des... Verhiiltnisse. welches zwischen den Aussagen der theoretischen Physik
und der Wirklichkeit besieht.’
32The notable difference from Einstein’s treatment is that Weyl, under the influence of
Levi-Civita, has already introduced the notion of a “geodetic” coordinate system (p. 100 ff.) in
which the components of any (contravariant, covariant) vector are unaltered by an “infinitesimal
parallel displacement” (infinitesimalen Parullelverschiebung); the transformation law of this parallel
displacement (involving the Christoffel symbols of the second kind) is then used to briefly
characterize a “geodetic vector field”, which in the 4th and 5th editions of Raum-Zeit Materie
(op.&., note 26) will be baptized as the “affine connection” (also, “guiding field”, Fiihrungsfeld) of
the manifold. Weyl goes on to express the Riemann curvature tensor (p. 108) in terms of the
three-index Christoffel symbol (of the 2nd kind), i.e. the components of the affine connection.
Weyl, Reichenbach and the Epistemology of Geometry 843

“purely infinitesimal”, that Einstein had achieved only a partial victory over
ferngeometrisch or fernvorstellung prejudices and the hoary old metaphysical
concept of substance.
Levi-Civita’s paper, though purely mathematical, was a first product of
the new attention Einstein’s theory had brought to differential geometry;
Levi-Civita himself had, together with his teacher, Giorgio Ricci, published
the first systematic treatment of the tensor calculus in a paper in 1901.33 In
his 1917 paper,34 Levi-Civita showed how the fundamental notion of
Riemannian curvature could be non-metrically expressed in terms of the
concept of an infinitesimal displacement of a vector parallel to itself.35 Weyl
generalized Levi-Civita’s result, removing an unnecessary embedding restric-
tion, and in two papers, published still in early 1918, laid the framework for
his unified theory. In the first of these, he sketched his unified theory which
incorporates electromagnetism into the metric of space-time; in the second he
provided a mathematical exposition of what he termed a “pure infinitesimal
geometry” which removed, as he put it, the last ferngeometrisch
“inconsistency”, a remnant of the Euclidean past, from Riemann’s theory of
manifolds.36 From now on, Weyl’s position will be that ‘A genuine Local-
Geometry (wahrhafte Nahe-Geometrie) can only be acquainted with a prin-
ciple of transport of length from one point to another infinitely adjacent to
it’; in other words, Riemann’s geometry goes only half-way towards the ideal
of a pure infinitesimal geometry since it assumes a path-independent, distant
comparison of line elements.37 In Weyl’s new geometry, both direction (like
Riemann) and length (unlike Riemann) are not, in general, integrable. In
such a Weyl space, as it is now known, lengths at one and the same point
alone can be compared but not from one point to another lying at a finite
distance. The epistemological attraction for Weyl of a purely infinitesimal
s3‘MOthodes de calcul differential absolu et leurs applications’, in Mathematische Annaien,
Vol. 54, pp. 125-201. This paper was an important resource for Einstein (and Grossmann) in the
mathematical development of general relativity.
“‘Nozioni di parallelism0 inuna varieta qualunquee conseguente specificazione geometrica della
Curvatura Riemanniana’, in Rendiconti de1 Circolo di Matematica di Palermo 42 (1917).
351f a (contravariant here, but alternately covariant) vector A’ at the points‘xk is parallelly
displaced to an indefinitely adjacent point xk + dx,, then the resulting vector A’ + 6A’ is determined
by the expression

6A’ = - {‘,k} Xdx,, ({j,k} = 1-;, = I-‘,)

where the 40 independent magnitudes Y,, (equivalently, Christoffel symbols of the second kind) are
expressed in accordance with Riemannian geometry by the metric tensor and its first derivatives: for
a clear discussion, see L. Silberstein, The Theory of Relativity, 2nd edition (London, Macmillan,
1924), pp. 346ff. In Levi-Civita’s paper, a construction is given in which the manifold where
parallel-transport is defined is embedded in a ten-dimensional flat Euclidean space.
36‘Gravitation und Elektrititlt’, in Sitzungsberichte. Preussischen Akademie der Wissen-
schaften, Physikalische-Mathematische Klasse (1918) pp. 465480; ‘Reine Infinitesimalgeometrie’,
Malhematische Zeitschrift 2 (1918), pp. 384411, p. 385.
37‘Gravitation und Elektrizitat’ (ibid.), p. 466; Raum-Zeit6Materie, 3rd Edn (Berlin, J. Springer,
1919), p. 91; Space-Time-Matter (op.&., note 26) p. 102.
844 Studies in History and Philosophy of Science

foundation for differential geometry was direct: it was fully d’accord with the
principle that understanding nature stemmed from formulating its behavior
in the infinitely small.
Since there is no longer an invariant unit of length, in order to have a metric
at all in a Weyl space, it is necessary first to specify an arbitrary unit length at
each point, that is, to “gauge” (umeichen) the space; mathematically, this is
accomplished by adding to the structure of the Weyl space a “pseudo-vector”
field which assigns to each point of the space a unit vector. This essentially
means that the choice of a local coordinate system also involves the choice of
a local unit of length. Compatibility of the affine structure developed from the
notion of infinitesimal vector parallel transport with the metric requires
permitting a special choice of gauge (“geodetic gauge”) for the infinitesimal
neighborhood of a point P such that vector length I is unaltered in parallel
displacement from P. But in general, a vector’s length changes under transport
from P(x’) to P’(x’ + dx) where the standard of length at P is carried over.
This is given by a similarity transformation,
dl = -ldy, (1)
where 9 is a function of position independent of the displaced distance.
However, the arbitrarily chosen gauge at P may be altered according to
I’ = 2
where the gauge factor A is an always positive continuous function of the
coordinates. In this case, in place of (l), we have
dl’ = - Pd# (1’)
where
dyl’ = dq - d;llA.
Weyl shows that if the desired condition that dl vanishes at P is met, then
dq = C p,dx’
which is to say that dy is a linear diffeiential form.
Characterization of the metric therefore requires two “fundamental forms”,
the quadratic differential form of Riemannian geometry,
ds = gikdxidXk (Einstein summation convention)
and the linear one just defined. These are defined up to the “gauge transfor-
mations”
ds’ = ids and
d# = dy, - d log il (d log A = dAl2)
Weyl would thus demand that natural laws not only satisfy the Einstein
condition of general covariance, i.e. invariance under arbitrary (but sufficiently
continuous) transformations of the coordinates, but also the more restrictive
Weyl, Reichenbach and the Epistemology of Geometry 845

condition of “gauge invariance”.38 By the vector operation “curl” (rot), Weyl


then defines from v, an antisymmetric tensor, which provides a measure of the
extent of the non-integrability at each point (that is, the departure from
Riemannian geometry there). In components it is written

where q is the function already noted, with the consequence that when Fjk=O,
the length of a vector is integrable (path independent), i.e. the special case of
Riemannian geometry is recovered.
Now Weyl’s discovery was that this tensor is formally identical to the
electromagnetic field tensor (the so-called Faraday tensor) and thus to the first
system of Maxwell’s equations if q is seen as the potential of the electromagnetic
field. It is therefore only natural to identify, as Weyl immediately did, these two
structures. The startling meaning of this identification is that all the phenomena
of nature are represented in the 14 independent coefficients pi and g, of the now
expanded metric; these in turn are to be deduced from a single universal
“world-law” (i.e. an action law) of the highest mathematical simplicity.39 In
‘*Space-Time-Mutter (op. cit., note 26), p. 286; “more restrictive” since Weyl now demands that
laws of nature be of “gauge weight” 0. This additional invariance means that the general solution
of the field equations now contain 5 arbitrary functions (instead of 4, as in Einstein’s theory), hence
that there be 5 identities (“superfluous equations”) between the 14 field equations. In accordance
with the theorem of integral invariants (Noether theorem), according to which a conserved quantity
corresponds to each invariant of a Lagrangian system, Weyl demonstrates, in an extremely
complicated discussion, that just as conservation of energy-momentum corresponds to coordinate
invariance, “conservation of electricity” (both the electric potential and the current satisfy the
boundary condition %++/%i=O) corresponds to gauge invariance, as is reflected in the fact that the
Maxwell equations are of gauge weight 0. See Raum-Zeit-Materie 5th Aufage (op.cit., note 26),
pp. 3088317. Weyl takes this demonstration to have provided ‘a strong formal argument in favor
of the theory: as origin of the conservation law of electric charge a priori must be expected one new,
single arbitrary function involving an invariance property (Invarianzeigenschuff) of the field laws.’
Geometrie und Physik op.&. , note 1 l), p. 54.
“‘Reine Infinitesimalgeometrie’ (op.cit., note 36) p. 385, footnote 4: ‘I am audacious enough to
believe that the totality of physical phenomena permit deduction out of a single universal world-law
( Weltgesetz) of highest mathematical simplicity.’ In a mode of procedure very much in the style of
the Mie-Hilbert theory of matter, Weyl would then seek in the next few years, without success, to
find a single action principle expressing the force law governing the interaction between the g,, and
the pi; this is to have the form of an integral invariant (Lagrangian density), additively combining
a gravitational and an electromagnetic component; see Space-Time-Mutter, Sections 35536 and the
later. somewhat chastened, abbreviated approach in Raum-Zeit-Materie 5th AuJage (op. cit., note
26), Section 40. This failure to univocally determine a “world-function” played no small part in the
critical reception of Weyl’s theory. Given the widespread use of Lagrangian and Hamiltonian
methods today, it is instructive to observe that Pauli (ap.cit., note 54 below), already in 1921,
expressed skepticism towards what he saw as the too little empirically oriented “Gottingen
approach” of seeking such a Weltfunktion: ‘it is not at all self-evident from a physical point of view,
that physical laws should be derivable from an action-principle. It would, on the contrary, seem far
more natural to derive the physical laws from purely physical requirements, as was done in
Einstein’s theory...’ (p. 201). Some years later, Eddington expressed a similar opinion in the face of
the difficulties posed by quantum mechanics: ‘Least action seems to fail here because it attempts to
jump by a formal device a territory comprising the problems of electron structure and quantum
theory requiring other conceptions than those of field physics. ’ ‘Universe: Electromagnetic-
Gravitational Schemes’, Encyclopedia Britannica, 13th Edn (1926), pp. 907-908. For a critical
assessment of Weyl’s failure to find a suitable Weltfunktion, see P. Bergmann, Introduction to the
Theory of Relativity (New York, Dover, 1976; first published in 1943), ch. 16.
846 Studies in History and Philosophy of Science

setting out to remove what Weyl saw as a glaring Schiinheitsfehler in Riemann’s


theory of manifolds, Weyl had produced a geometric theory unifying electricity
and gravitation. Since these were the only physical forces recognized in
1918, Weyl was led to triumphantly proclaim the unity of geometry and
physics:

Everything real (Wirkliche) that transpires in the world is a manifestation of the


world-metric: Physical concepts are none other than those of geometry.40

Such declarations elicited no small amount of criticism from Einstein and


others; despite their stridency, what Weyl seems to have meant is simply that it
is the task of geometry to investigate the “essence of metric concepts”, while
physics is concerned to determine “the law” (i.e. the integral invariant
formulated as an action principle) according to which the actual world is
singled out from all other possible four-dimensional metric spaces.41
Now for Weyl, the essence of the concept of metric lies in the concept of
congruence, which is a “purely infinitesimal” concept;42 here the metric is
viewed solely as a structure of the continuous field whereupon the concepts of
vector and “tract” (M-e&e) employed for its characterization have in them-
selves “nothing to do” with material measuring rods.43 Within the epistemology
of a purely “contiguous physics” and of “purely infinitesimal geometry”, Weyl
distinguishes the “nature” of the metric, which is a priori and the same at each
point of space, from its “orientation” from one point to another indefinitely
nearby, this being variable and pace Riemann and Einstein, locally dependent
upon the distribution of matter.
As a corollary to these epistemological foundations of exact science, Weyl
will take up the so-called “Raumproblem” whose Helmholtz-Lie solution,
based as it is on the ferngeometrisch postulate of free mobility, is neither
adequate to the non-homogeneously curved spaces permitted by Riemann’s
theory of manifolds nor survives the scruples of Weyl’s stricter “purely
infinitesimal” point of view. In so doing, he locates the essence of the metric,
hence of congruence, in the idea of the infinitesimal orthogonal group, and he
is able to prove, under the assumption that the metric uniquely determines the
affine connection, that the group of infinitesimal rotations at a point (for n
dimensions) that can be characterized as the set of linear transformations
leaving a non-degenerate quadratic differential form invariant is the only group

“‘Reine Infinitesimalgeometrie’ (op.cit., note 36) p. 385.


411bid.: ‘The sole distinction between geometry and physics is this: that geometry investigates
generally what lies in the essence of metric concepts, while physics determines the law through
which the actual world is singled out from among all possible four-dimensional metric spaces of
geometry and explores its consequences.’
@‘Die Einzigartigkeit der Pythagoreischen Massbestimmung’ (op.cit., note 23) pp. 114115.
43‘Ueber die physikalischen Grundlagen der erweiterte Relativitatstheorie’ (op. cit., note 66
below), p. 473.
Weyl, Reichenbach and the Epistemology of Geometry 847

whose mappings of a vector onto itself are “volume true”, i.e. accord with the
concept of congruence. This demonstrates the “uniqueness of the Pythagorean
determination of measure”, hence its a prioricity and provides a new solution
to the “Raumproblem”.44 For our purposes, I wish here only to call attention to
the thoroughgoing character of Weyl’s, naturphilosophische perspective, which
unites physics and geometry; and thus to show how deeply rooted is his
philosophical opposition to what will become the standard philosophical view
of measurement in GTR.
Beginning with the third edition of Raum-Zeit-Materie, published in 1919,
Weyl also presents a development of Einstein’s theory from the perspective of
his more general geometry. It is worth spending a few moments here to sketch
this mathematical characterization of general relativity, both because it creates
the mathematical machinery for other attempts (including Einstein’s own,
which occupied him to the end of his life in 1955) to expand general relativity
into a unified theory of fields, and because it has become more or less standard
in contemporary accounts of general relativity. 45 In Einstein’s original treat-
ment, general relativity is developed by assuming Riemannian geometry at the
outset. But Weyl now develops the differential geometry on which general
relativity is based in three stages.
At the most fundamental level is the concept of a topological manifold,
corresponding (as in Riemann) to the notion of an empty world.46 To such a
manifold is added an affine connection based, as in Levi-Civita, on the
notion of infinitesimal parallel vector shift; physically, an affine space is
interpreted as a world endowed with a gravitation-inertial field, to which
Weyl gives the suggestive name, Fiihrungsfeld (“guiding field”) and in
which affine geodesics are instantiated as the inertial trajectories of
uncharged ‘test particles’, and gravitational “force” as occasioning (slight)
departures from these inertial trajectories. Significantly, now the components
of the affine connection (the 40 quantities I;J, and not those of the metric
tensor g,, are taken as the field strengths of the gravitational field; indeed,
this produces the only consistent way in which to view the Principle of
Equivalence (since the I:, is not a tensor, and hence can be locally

44‘Die Einzigartigkeit der Pythagoreischen Massbestimmung’ (qcit., note 23); Mathematische


Analyse des Raumproblems (note 23), lectures 7 and 8. Weyl uses a method that combines Lie’s
theory of continuous groups with infinitely small “virtual” variations of the mappings onto itself of
an n-dimensional vector body radiating from a point. For a recent discussion of Weyl’s solution,
see the paper by Erhard Scheibe in W. Deppert et al. (eds), Exact Sciences and their Philosophical
Foundations (op.cit., note 49 below).
45A fundamental text in this regard is E. Schriidinger, Space-Time Structure (Cambridge,
Cambridge University Press, 1950).
‘?n ‘Reine Infinitesimalgeometrie’ (op. cit., note 36), p. 286, Weyl refers to a topological manifold
as an “empty world”, sometimes also referring to it as an “amorphous continuum”, but such a
manifold is not to be viewed as structureless, since “the ‘topological skeleton’ determines the
connectivity of the manifold in the large.” Philosophie der Mathematik and Naturwissenschaft,
pp. 63 and 66 (op.cit., note 66 below).
848 Studies in History and Philosophy of Science

“transformed away”). Finally comes the notion of a metric space in which it


is possible to speak of the distance (respectively, time interval) between two
point-events. This is a world endowed with a causal structure and either (as
in Einstein) a path-independent transport of length, or, with Weyl, only a
path-dependent one.
A crucial consideration is that the metrical structure be compatible with
the affine structure so that affine geodesics are also geodesics of the metrical
space. Weyl proved that the metric univocally determines the affine connec-
tion; this relation of compatibility he calls “the basic facts (Grundtatsache) of
infinitesimal geometry”.47 Hence, even though the notion of metric is logi-
cally the more fundamental conception, what is most significant about this
treatment is that it shows how Einstein’s unification of gravitation and
inertia (from the Principle of Equivalence) can be conceived solely as the
structure of the affine connection of the manifold of space-time. This
amounts to, as it were, “peeling away” the metrical structure of the manifold
to reveal the structure responsible for the effects of inertia and gravitation
(hence Weyl’s term, Fiihrungsfeld, “guiding field”).48 This approach was
subsequently adopted by Einstein and has become more or less standard.
What is not standard is that, for Weyl, there are two different kinds of
curvature of space-time, one corresponding to the gravitational-inertial field,
the other corresponding to the electromagnetic field, which also enters into
consideration as part of the metrical structure.49

3.

When, in early March, 1918, Weyl first informed Einstein of his result,
Einstein wrote back that he thought it a “a stroke of genius of the first
rank”.50 But within a few days, Einstein was to formulate an objection that, as

47Raum-Zeit&A4aterie, 3rd Edn (op.&., note 37) p.111; see the discussion at the end of Section
2 below and the Appendix to this paper.
“sCoffa (op.&., note 18) p. 279, uses this unstripping metaphor.
491n addition to providing the basis of the modern mathematical treatment of GTR, Weyl also
proposed a method for solving the problem of measurement in GTR (see below), made
fundamental contributions to formulating the equations of motion, discovered a large class of exact
solutions to the Einstein-Maxwell field equations, and helped to create relativistic cosmology.
These contributions are surveyed in .I. Ehlers, ‘Hermann Weyl’s Contributions to General
Relativity’, in W. Deppert et al. (eds), Exact Sciences and their Philosophical Foundations,.
Proceedings of the Hermann Weyl Congress, Kiel, 1985 (Frankfurt am Main-Bern. Verlag Peter
Lang, 1988) pp. 833105. In Ehlers’ assessment, ‘Apart from Albert Einstein, nobody has
contributed more to the conceptual clarification of the general theory of relativity than Hermann
Weyl.’ (See pp. 8485.)
5”In a postcard to Weyl dated 6 April 1918; quoted in Straumann (op.&., note 25), p. 415.
Eddington (op.cit., note 52 below), regarded Weyl’s theory as “unquestionably the greatest advance
in relativity after Einstein’s work” (p.198). Of course, Weyl’s work was the spur for Eddington’s
own efforts toward unification.
Weyl, Reichenbach and the Epistemology of Geometry 849

we shall see, proved to be decisive (to which we will come in a moment).


Nonetheless Weyl’s paper was communicated by Einstein to the Prussian
Academy of Sciences (after some negotiation by Einstein on Weyl’s behalf with
Walter Nernst, who presumably didn’t understand it) and it was published,
with an appended note by Einstein outlining his objection (which Nernst had
insisted on) together with a reply by Weyl, in the proceedings of the Academy
in May 191LL51
Weyl’s theory, put forward in two explanatory versions, received an
unusual reception from the theoretical physics community, beginning with
Einstein himself. This response still awaits an adequate historical treatment;
and it seems to tell much about authority and how challenges to authority
are met within science. There is, above all, the manner of Einstein’s reaction.
The received view is that Einstein simply pointed out that Weyl’s theory was
not in accord with the empirical facts and that was the end of the matter.
But Einstein’s response-over the next six or so years!-was considerably
more ambiguous and the actual sequence of events is considerably more
complicated.
It is of course true that Einstein expressed this objection privately in
correspondence with Weyl and then publicly, in the aforementioned note
appended to Weyl’s Prussian Academy paper, and then again indirectly in the
well-known “popular” lecture entitled “Geometry and Experience” given in
January, 1921, to the Prussian Academy. This objection was taken up,
amplified, and broadcast to the wider physics community by Wolfgang Pauli jr.,
in the canonical encyclopedia article on relativity he wrote in 1920 (when he was
20!). The objection seemed to show that Weyl’s theory was not in agreement
with observation.
In fact, in its original version, Weyl’s theory predicted effects that, as
Eddington showed, were below the threshold of observations2 and, as it later
turned out, within the limits of quantum mechanical uncertainties.53 As Pauli
himself had helped to show in one of his first scientific publications, Weyl’s
theory, though containing field equations of the fourth (and not the second)
order, yielded Einstein’s field equations as a special case and agreed on two

“‘Gravitation und Elektrizitat’ (op. cit., note 36); for Nernst’s obstruction, see Sigurdsson (op. cit.,
note 30), pp. 166167.
“The Mathematical Theory of Relativity (Cambridge, Cambridge University Press, 1923) p. 207.
After making some order of magnitude assumptions which set Fik comparable to the force at the
surface of an electron, Eddington concludes, ‘Thus dN1 [the ratio of the change in length of an
infinitesimally transported vector to its original length at P-TR] would be far below the limits of
experimental detection.’ Pauli’s text is cited in note 55 below.
53For example, a standard text reports, ‘The strength of Einstein’s objection seems not
as powerful now as at the time when it was raised, since we know the classical physics does
not describe atomic phenomena without certain quantum-theoretical modifications.’ R. Adler,
M. Hazin and M. Schiffer. Introduction to General Relativity, 2nd Edn (New York, McGraw-Hill,
1975) p. 506.
850 Studies in History and Philosophy of Science

predictions derivable from Einstein’s theory: the gravitational red shift and the
advance of the perihelion of Mercury.54
What then was Einstein’s objection? We do observe that measuring rods
retain their length under transport in electromagnetic fields; prima facie, this is
evidence that Riemann’s geometry, not Weyl’s, is the geometry of space-time.
More substantively, Einstein argued as follows: if Weyl’s theory is correct, then
the spectral lines emitted by atoms would not be very sharp and well-defined as
in fact they are observed to be. For if two atoms, say of hydrogen, are once
together in one space-time region and then transported via different paths to
another region of space-time where they are brought together again, then,
according to Weyl’s theory, we should generally observe a difference in their
spectral lines corresponding to their past histories, i.e. to the differing values of
the electromagnetic field strengths in the space-time regions they occupied in
the interim. Pauli, indeed, showed that no matter how small the initial
difference in the spectral lines of the two atoms posited by Weyl’s theory, this
difference would “increase indefinitely in the course of time”.55 But astronomi-
cal observation tells us that hydrogen atoms everywhere in the heavens exhibit
the same spectral signature. So Weyl’s theory did not, apparently, correspond
to the facts of observation.
At the time, this objection was largely taken as decisive by the community of
physicists, though, as has been noted, perhaps it should not have been. In no
small measure, this is of course due to the enormous prestige of Einstein as the
‘%ee ‘Merkurperihelbewegung und Stralenablenkung in Weyls Gravitationstheorie’, Verhand-
lungen der Deutschen Physiknlischen Gesellschaf 21122 (1919), 742-750. In another early paper ‘Zur
Theorie der Gravitation und der Elektrizitlt von Hermann Weyl’, Physikalischen Zeitschrift 20
(1919), 457467, Pauli took up the issue of the action function, seeking in particular, a function
yielding static, spherically symmetric singularity-free solutions (corresponding to the atomic
composition of matter). Weyl did succeed in finding solutions containing singularities, i.e.
corresponding to an electrical particle in which the phase quantity a, (the electrostatic potential) has
a singularity at the radial center, which briefly led him to proclaim, “Matter is accordingly a true
singularity of the field.” (See Space-Time Matter, pp. 298-300.) However, the failure to find
singularity-free solutions, hence to resolve “the problem of matter”, coupled with the fact that the
differential equations of Weyl’s theory were so complicated it was not possible to integrate them,
were ultimately the stumbling blocks on which Weyl’s theory foundered. See the article by
V. Bargmann, ‘Relativity’, in M. Fierz and V. Weisskopf (eds), Theoretical Physics in the Tlventieth
Century; A Memorial Volume to Wovgang Pauli (New York, Interscience Publishers, 1960)
pp. 1877198.
“Taking the electrostatic potential p to be a function of time, Pauli derives, from Weyl’s gauge
transformations, the equation r = roe”‘@, where r is the proper time and a is a factor of
proportionality. Pauli explains this equation as follows: ‘Let two identical clocks C,, Cz, going at
the same rate, be placed at first at the points P,, at an electrostatic potential (o,. Let the clock C,,
then be taken to point Pz, at potential pa, for t seconds, and then finally returned to P,. The result
will be that the rate of clock C,, compared with that of clock C,, will be increased or decreased,
respectively, by a factor exp[-a((o, - q,)t] (depending on the sign of a and of (o, - p,). In
particular, this effect should be noticeable in the spectral lines of a given substance, and spectral
lines of definite frequencies could not exist at all. For, however small a is chosen, the differences
would increase indefinitely in the course of time, according to [this equation]‘. Theory of‘ Relativity
(New York, Pergamon Press, 1958) p.196; the original German text was published in 1921 as
‘Relativitltstheorie’ in the Encykloptidie der mathematischen Wissenschuften V19 (Leipzig, B.C.
Teubner).
Weyl, Reichenbach and the Epistemology of Geometry 851

discoverer of general relativity. In fact, and as hinted above, in several letters to


Weyl during 1918 and 1919, Einstein reveals that his opinion of Weyl’s theory
is far less one-sided than his public posture would suggest.56 And there can be
little question that recurring bouts of doubt as to the total invalidity of Weyl’s
approach, and of the related generalization offered by Eddington are evident.
For even as he repeatedly remarked during the period 1919-1925 that this route
led only to exasperating dead-ends, he continued to explore his own variants of
the Weyl/Eddington unification schemes in the initial steps of what was to be a
three-decade-long futile effort to construct a unified theory of fields.57 It is
almost as if Einstein could finally quiet his own nagging self-doubts only by
exhausting all the possibilities he deemed reasonable within what he called the
“Weyl-Eddington complex of ideas”.58
Despite his enormous respect for Einstein, Weyl was not persuaded by the
Einstein objection. Instead, he adopted a two-pronged argumentative strategy
to counter it, in effect, producing a second explanatory version of his theory. On
the one hand, he provocatively responded that the behavior of physical objects
such as rods and clocks or, for that matter, atoms, has “as such nothing to do”
with the ideal metric notions defined by vector transference:

The functioning of these instruments of measurement is however a physical occur-


rence whose course is determined through laws of nature and which has as such

56Excerpts from letters in 1918 and 1919 are quoted in Straumann (op.&., note 24).
Banesh Hoffman also speaks of “Einstein’s official argument against Weyl’s theory”; Albert
Einstein; Creator and Rebel (New York, Viking, 1972), p. 223. The “unofficial” side of Einstein’s
objections to Weyl emerge clearly in his correspondence with M. Besso, his friend in Zurich
who was also a friend of Weyl and favorably disposed to the logic of Weyl’s theory; see his
letters to Besso of 4 December and 12 December 1918 and of 26 July 1919 in P. Speziali (ed.)
(note 58 below). F. Herneck has uncovered a recording Einstein made in Vienna on 14 January
1921, in which Einstein’s warning, that unless the interval element ds is connected with “the
observable facts”, a “reality-strange” ( Wirklichkeitsfremden) theory is the result, is taken
as an indirect criticism of Weyl; Einstein und seine Welrbild (Berlin, Der Morgen, 1976),
pp. 103-108.
s71n addition to the paper of 1919 (op.&. , note 68 below), Einstein’s engagement can be
tracked through a large number of papers in the 1920s in the Sitzungsberichte of the Preussichen
Akademie der Wissenschaften, Physikalische-Mathematische Klasse: ‘Ueber eine naheliegende
Ergiinzung des Fundaments der allgemeinen Relativitiitstheorie’ (1921) pp. 261-264; “Zur
all,emeinen RelativitSitstheorie” (1923), pp. 32-38 and 7677; ‘Zzrr afien Feldtheorie’ (1923),
pp- 1377140; ‘Einheitliche Feldtheorie v& Gravitation und Elektriz& (1925), pp. 414419, as
well as in ‘The Theory of the Affine Field’, Nature 112 (1923), 448449 and in ‘Eddingtons
Theorie und Hamiltonisches Prinzip’, an Anhang to the German translation of Eddington’s
work (op. cit., note 52), Relativitatstheorie in M&hematischer Behandlung (Berlin, J. Springer,
1925), pp. 367-371. The conclusion there, essentially no different from similar verdicts
expressed since 1919, reveals the continuing attraction of the Weyl-Eddington approach: ‘For
me, the end result of this consideration unfortunately consists of the impression that the
Weyl-Eddington deepening of geometric foundations is capable of bringing us no progress
in physical knowledge; hopefully, future development will show that this pessimistic opinion
has been unjustified.’ As we shall see, Reichenbach will take to heart the first clause of this
opinion.
‘“Letter to Besso 25 December 1925, P. Speziali (ed.), Albert Einstein Michele Besso Correspon-
dance, 1903S1955 (Paris, Hermann, 1972), p. 215.
852 Studies in History and Philosophy of Science

nothing to do with the ideal process of congruent transplantation of world intervals


(Verpjlanzung von Weltstrecken).59

For Pauli, this meant that although there was no longer a “direct contradic-
tion with experience”, there also no longer existed an “immediate connection”
between electromagnetic phenomena and the behavior of measuring rods and
clocks; hence, the connection between electromagnetism and the world metric
posited in Weyl’s theory becomes “purely formal”. Eddington similarly inter-
prets Weyl as thus giving up any claim to characterize the geometry of the real
world, by providing only a “graphical representation”, i.e. a kind of conven-
tional representation, of “world geometry”.60 After all, if geometrical relations
are not concerned with measuring rods and clocks, what are they concerned
with?61
Weyl’s point, however, though Pauli’s response does not directly address it,
was that such a response ducks an explanatory burden that is essential from the
viewpoint of a consistent theory of fields or a systematic theory of field-matter
interactions.62 For it is just wrong-headed or, as Weyl expressly states,
“perverse” to use physical bodies such a rods and clocks, which are indicators
of the gravitational field, as at the same time instruments to stipulate metric
relations. To do so is just to treat as a definition (“rigid rod”, “clock”) what
should be explained, i.e. should be derivable from a systematic theory.63 In terms
of such a systematic theory, “Einstein’s dejinition of measure determinations in
the metrical field with the help of measuring rods and clocks has validity only
as a preliminary connection to experience just as does the definition of electrical
field strengths as the ponderomotive force on a unit charge”. In order, in Weyl’s
terms “to close the circle”, it is

necessary, once a suitable action law has been set up, to prove that here, the charged
body under the influence of the electric magnetic field, there, the measuring rod under
the influence of the metrical field, exhibit, as consequences of the action laws, that
59‘Eine Neue Erweiterung der Relativit%tstheorie’, Annalen der Physik 59 (1919), 101-133, at
p. 113.
“‘The Mathematical Theory of Relativity (op.cit., note 52), Section 83 (‘Natural geometry and
World geometry’). ‘The new view entirely alters the status of Weyl’s theory. Indeed it is no longer
a hypothesis, but a graphical representation of the facts, and its value lies in the insight suggested
by this graphical representation’. The antithesis of “natural geometry” versus “graphical represen-
tation” will recur in Reichenbach’s examination of Weyl’s theory (see below). Pauli subsequently
also finds this interpretation to characterize the situation very well; see his review of the German
translation of Eddington’s book in Die Naturwissenschaften 13 (1926), 273-274 and also his letter
to Eddington of 20 September 1923, thanking Eddington for sending his book; A. Hermann et al.
(eds) Wolfgang Pauli Wissenschaftlicher Briefwechsel, Bd. 1 1919-1929 (New York/Heidelberg,
Berlin, Springer, 1979), pp. 115-l 19.
“As Coffa (opcit., note 18), pp. 283-284, forcefully puts this objection.
62Weyl changes his own interpretation of his theory along this axis; see ‘Feld und Materie’
(op.cit., note 66 below) and the 5th edition of Raum-Zeit-Materie (op.cit., note 26), Section 38.
63‘Eine neue Erweiterung der Relativitltstheorie’ (op.cit., note .59), p. 113: ‘If then these
instruments [i.e. measuring rods and clocks] also play an unavoidable role as indicators of the
metric field...then it is apparently perverse (verkehrt) to define the metric field through indications
taken directly from them.’ As we shall see, Reichenbach’s discussion concludes just the opposite.
Weyl, Reichenbach and the Epistemology of Geometry 853

behavior we had originally utilized for the physical definition of the field
magnitudes.G4

General relativity, in the conception of Einstein, is not, of course, systematic


in this sense, relying as it does on the notion of a “practically rigid rod” that
corresponds to “congruence at a distance” (i.e. path independent transport of
length). However, from the perspective of Weyl’s elegant generalization of
Riemannian geometry, the fundamental metric concept, congruence, is only to
be properly conceived as a “purely infinitesimal” concepV5 to do so renders
Einstein’s “practically rigid rod” an unprincipled and gratuitous assumption.
As we shall see in a moment, this assumption is “built into” Reichenbach’s neo-
conventionalist account of the geometry of space and time, with an argument
that has Weyl’s alternative squarely in its sights.
Clearly, Weyl had to account for why we do observe the congruence-
preserving behaviors of rods and clocks that we do, as well as for the constancy
of spectral lines of atoms. He does so (and this is the other component of Weyl’s
explanatory strategy) by invoking a dualism regarding the manner in which
physical quantities are determined, a distinction which he sees as reprieving his
theory from the empirical refutation sketched by Einstein and elaborated by
Pauli.
Physical quantities are fixed either by a body’s following a “tendency of
persistence” (Beharrungtendenz) or by its “adjustment” (Einstellung) to the
field strengths where it is, a distinction made concrete by appealing to the
different physical behaviors of a spinning top and the magnetic needle of a
compass.66 Whatever its initial orientation, the axial direction of a spinning
top is transferred from instant to instant by a tendency of persistence, that is
to say, it is governed by the inertial (a.k.a. “guiding”) field. On the other
hand, the magnetic needle of a compass adjusts to the value of the magnetic
field wherever the compass is carried; “adjustment....enforces a definite value
that is independent of past history and hence reasserts itself after any

“Raum-Zeit-Materie, 5 Aufage (op.cit., note 26), p. 298 (from a section new to this edition).
“Mathematische Analyse des Raumproblems (op.cit., note 23), p. 47: “Die Metrik hangt am
Begriffe der Kongruenz, der jedoch rein infinitesimal gefasst werden muss.” (Original emphasis.)
66This distinction seems to have been made in print first in Weyl’s report ‘Elektrizitlt und
Gravitation’, Physikulische Zeitschrift 21 (1920), 6491651, at p. 650, to the infamous 86th gathering
of German natural scientists (19-25 Sentember 1920) at Bad Nauheim; the distinction is more fully
discussed in ‘Electricity and Gravitation’, Nature 1Oi (1921), 8OC802; in ‘Ueber die physikalischen
Grundlagen der erweiterten Relativitltstheorie’, Physikalische Zeitschrift 22 (1921) 473480; in
‘Feld und Materiel, Annalen der Physik 65 (1921) 541-563; and in Weyl’s own report on what
transpired at Bad Nauheim, ‘Die Relativitltstheorie auf der Naturforscherversammlung in Bad
Nauheim’, Jahresbericht der Deutschen Mathematische-Vereinigung (1922), pp. 51-63. It is remark-
able that his distinction is still retained some thirty years later, long after Weyl abandons his own
unified theory, in the form of an objection raised from the point of view of a (now future)
“systematic theory” to the rigid rods and clocks of Einstein’s theory; the objection occurs in an
appendix added to the 1949 English translation of his 1927 philosophical monograph, Philosophie
der Mathematik und Naturwissenschaft (Miinchen/Berlin, R. Oldenbourg); see Philosophy of
Mathematics and Natural Science (Princeton, Princeton University Press), Appendix F, p. 288.
854 Studies in History and Philosophy of Science

disturbances and any lapse of time as soon as the old conditions are
restored”.e’
From Weyl’s new perspective, the Einstein-Pauli “prehistory” objection to
his theory presupposes that measuring rod lengths and clock periods are altered
(or not) through a time-dependent process of persistence, whereby the magni-
tude in question at a given instant is some function of its magnitude at a
previous instant. But given the distinction Weyl has called attention to, this is
not at all a necessary presupposition, in which case the explanatory burden runs
in the other direction. If, for instance, a measuring rod is moved around within
a physical field assumed to be inhomogeneous, i.e. where the phase quantities
have different magnitudes at different points, an account is surely required as to
why we detect no variable behavior in our measuring rod. One prima facie
reason may be that such “deforming” forces at each space-time point are
present but are counteracted by electromagnetic forces within the atom which
thus produce a state of force equilibrium.
To motivate this conception, Weyl here again invokes the analogy of the
“ponderomotive force” exercised by an electrical field on a charged body within
the field, and to the fact that the charged body itself interacts with and changes
the field strengths where it is located. As Weyl observed, the character of such
interactions are normally non-linear, yet the unit of negative electrical charge
represents one equilibrium state. The situation is in principle no different in the
case of a massive physical object located in a gravitational field, though the
force of gravitation is many factors of magnitude weaker than electromagnetic
forces. Just as the charged body “carries along” a determinate value of charge
through its interaction with the electrical field, a massive object, such as a
measuring rod, carries with it a determinate magnitude representing the inter-
action of the gravitational forces of the field and the electromagnetic forces
obtaining between the rod’s constituent molecules and atoms. Hence a measur-
ing rod “adjusts” to the field strengths where it is now; it does not exhibit what
was for Weyl a qualitatively distinct, but miniscule (i.e. far below the threshold
of observation) quantitative difference in pattern of behavior, the tendency for
its length to persist, that is, to remain the length it was a moment ago at another
point of space-time. The common-sensical objection that of course the length of
a measuring rod persists in moving it from one end of a room to another can,
persumably, be countered by taking into account the crudity of our everyday
experience with middle-sized objects, which are not presumed to interact with
the “empty space” in which they (and we!) are located.
What is this field strength? Weyl appears to have been motivated by
Einstein’s 1919 “scalar-free field equations” (which were intended to provide a
more principled footing for the nefarious cosmological constant), to take the
relevant field strength here to be the equivalent in his theory to the Riemann
“7Philosophy ofMathematics and Natural Science (ibid.), p. 288.
Weyl, Reichenbach and the Epistemology of Geometry 855

scalar of curvature in Einstein’s, which has now become a constant, playing the
role of keeping electrodynamical forces within electrical corpuscles in equilib-
rium.68 Analogously to Einstein, Weyl can appeal to inertial-gravitational
forces as responsible for maintaining an equilibrium of intra-atomic electrical
forces. Hence, we observe that measuring rods display the behavior we famil-
iarly think they do: their length4ue to “adjustment”, i.e. constant force
equilibrium-is unaltered under transport in the gravitational field.e9 Weyl’s
response to Einstein and Pauli is, then, to say that the alleged “empirical
refutation” (prehistory objection) does not touch his theory, since the constan-
ties of “atomic clocks”, as also the congruence behaviors of measuring rods, are
to be accounted for as arising through Einstellung (as indeed they must in a
principled “systematic” theory), not through Beharrung, as the “prehistory
objection” wrongly presupposes.70

68The motivation is briefly hinted at in ‘Ueber die physikalischen Grundlagen der erweiterten
RelativitPtstheorie’, (op.cit., note 66), p. 414. In his ‘Spielen Gravitationsfelderim Aufbau der
materiellen Elementarteilchen eine wesentliche Rolle?‘, Silzungsberichte. Preussische Akademie der
Wissenschaften. Physikalische-Malhematische Klasse (1919), pp. 349-356, Einstein attempts to
reinterpret the cosmological term I, appended to his field equations in 1917 in an effort to satisfy
what he would come to call “Mach’s principle”, as a constant of integration, rather than as
hitherto, “a peculiar universal constant”. To do so, Einstein slightly modifies his field equations
without the cosmological term in a manner which explicitly accords gravitational force a role in
holding atoms together. Now the proportionality factor on the left-hand side is changed (from l/2
to l/4). But this modification causes the scalar R of the left-hand side (i.e. of R,,, - 1/4g,,R) to
vanish identically as does the scalar of the right-hand side (of ~ x 7”“) already when the
stress+energy tensor is written in terms of the Maxwell-Lorentz components of the electromagnetic
field and the divergence is taken. The modification enables him to derive the result that the scalar
of curvature R is constant (1) in all domains in which the current density of electricity vanishes
(“empty space”) and (2) on every world line of the motion of electricity (i.e. regarding electricity as
a moving charge density). Einstein then gives the following “intuitive” (anschaulich) interpretation
(p. 352): ‘The curvature scalar R plays the role of a negative pressure which outside of the electrical
corpuscles has a constant value R,. In the interior of each corpuscle there subsists a negative
pressure (R - R,), whose drop maintains the equilibrium of the electrodynamic forces. The
minimum of pressure (respectively, the maximum of the curvature scalar) does not alter wirh time
in the interior of the corpuscle.’ (Emphasis added-TR.) Finally, Einstein shows that in regions
where gravitational and electrical fields are present, 1=1/4&. A. Pais, ‘Subtle is the Lord...‘, The
Science and Life of Albert Einstein (New York, Oxford University Press, 1982), p. 257, terms this
paper, “Einstein’s first attempt at a unified field theory”.
@‘See Space-Time-Matter (op. cit., note 26), pp. 134K. for the definition of F (the scalar of
curvature in Weyl’s general metrical space) in terms analogous to that of the Riemann curvature
scalar R; on the basis of F, Weyl develops the notion of a “natural gauge”: it is to this “natural
gauge” that measuring rods and the frequencies of atomic clocks “adjust”. See also the discussion
in Raum-Ze&-Maferie, 5th Edn, Section 40, and p. 303: ‘measuring-rod lengths and frequencies of
atomic clocks are conserved on the basis of the natural gauge (natiirliche Eichung), thus in fact are
determined through adjustment (Einstellung) to the radius of curvature.’
“To my knowledge, the literature since the 1920s contains no account of this aspect of Weyl’s
defence of his theory. The physicist Arnold Sommerfeld seems to have thought, at this point, that
there was only a minimal difference between Einstein and Weyl, and he urged Einstein to meet Weyl
half-way; see his letter to Einstein of 10 August 1921: ‘I have the feeling as if between you and
[Weyl] there is only a really small distinction (ein ganz kleiner Unterschied). [Weyl] would overcome
the practical effect of his measuring rod alterations through his [concept of] “adjustment”
(Einstellung) and you would restrict the indeterminacy of the world function ( Weltfunktion)if you
would take the g,, as relational magnitudes.’ In Albert EinsteinlArnold Sommerfeld Briefwechsel,
A. Hermann (ed.) (BaseUStuttgart, Schwabe and Co., 1969), p. 87.
856 Studies in History and Philosophy of Science

Before we go on to Reichenbach, we need only see how the metric of


space-time is to be non-conventionally determined according to Weyl. In
Weyl’s theory, because of gauge invariance, it is not the g, themselves, but only
their ratios that have an empirically ascertainable meaning, a situation that, to
Einstein, robbed the interval ds of its empirical foundations.7l Moreover, if we
cannot, with Einstein and Reichenbach, stipulate that there are invariant
physical lengths and time intervals to which our mathematical notions are
“coordinated”, how is the metric to be empirically attested? Weyl’s answer is to
provide a so-called “geodetic method” involving two “directly observable”
physical processes, the propagation of light rays and the inertial trajectories of
mass points, both of which are required to fix the metric of a Weyl space.72 But
first comes a theorem, which Weyl published in 1921, according to which the
conformal and the projective properties of a metric space (in Weyl’s sense)
univocally determine its metric.73
This is a purely mathematical result, yet Weyl observes that the conformal
and projective properties have intuitively clear physical counterparts in the
theory of relativity. The conformal properties are given by the propagation of
light rays (“geodetic null lines”) which determine the causal structure of
space-time; hence, by observing the arrival of light at points in the immediate
neighborhood of a point 0, the ratios of the quantities g, at 0 may be
determined. But the propagation of light determines only the quadratic
differential form ds* (up to a conformal factor) while leaving the linear form v,
unrestricted. This latter may be fixed by considering that the projective
properties of the space correspond to the tendency of persistence of freely
moving point masses (inertial trajectories). If one assumes that the proper time
s may be read off from the motion of these unaccelerated ideal particles, then,
through comparison of two such point masses passing through 0 in different
directions, a unit of measure at 0 may be uniquely determined. The physical

“Letter to Weyl, 15 April 1918: ‘Lasst man den Zussamenhang des ds mit Massstab- und
Uhr-Messungen fallen, so verliert die Rel. Theorie iiberhaupt ihre empirische Basis’; quoted in
Straumann (op.cit., note 25), p. 416. This objection was publicly made by Einstein in September
1920 at the Bad Nauheim meeting; see the discussion of Weyl’s paper in Physikalische Zeitschrif
21 (1920), 651.
“Space-Time-Matter (op.cit., note 26), pp. 228-229 and Appendix 1, pp. 313-314. I am indebted
to David Malament for pointing out the unclarities of a previous treatment of this topic.
” ‘Zur Infinitesimalgeometrie: Einordnung der projektiven und der konformen Auffassung’, Kiinig-
lichen Gesellschaft der Wissenschaften zu Giittingen, Nachrichten. Math-Phys. Klasse (1921), pp.
99-112, esp. p. 100: ‘In the theory of relativity, projective and conformal properties have an
immediately intuitive significance.....(t)he tendency of persistence (Beharrungstendenz) of world
direction of a moving material particle, which forces upon it, if it is set loose in a determinate world
direction, a determinate “natural” motion, is a unity of inertia and gravitation...However, the
infinitesimal (light) cone [of null direction, given by d.? = O-TR] realizes in the neighborhood of
a world point the distinction between past and future; the conformal property is the action-
connection (Wirkungszusammenhang) of the world, through which is determined which world
points stand in possible causal connection with one another. Hence it is also a meaningful fact for
physics which comes to expression in the following theorem: Theorem I. The projective and
confbrmalproperties of a metric space univocally determine its metric.’ (The proof follows.)
Weyl. Reichenbach and the Epistemology of Geometry 857

significance of the theorem is then that the world metric can be fixed without
reliance on measuring rods and clocks, so that indeed, in itself, the notion of a
metric has nothing to do with measuring rods and clocks. It is worth noting that
this method still retains interest, 74 though there are problems with it, as there
are with any alternative.75

4.

Reichenbach’s initial views on the epistemology of geometry are expressed


. .
in his 1920 monograph Relatlvltatstheorie und Erkenntnis A Priori. Here
Reichenbach is concerned to show how the theory of relativity (GTR) has
necessitated a modzjication of Kant’s view of the a priori elements in knowledge;
in particular, only the sense of such elements as “constitutive of the object of
knowledge” is to be retained, while the sense of “apodictic certainty” is to be
jettisoned. Reichenbach has taken over Schlick’s definition of knowledge as a
“coordination” (Zuordnung) of concepts to “reality”, thus following Schlick
in rejecting altogether the Kantian problematic of the “schematism of the
categories” by denying any cognitive role at all to a faculty of intuition. In place
of Kant’s “analysis of reason”, Reichenbach proposes “the method of analysis
of science” (der wissenschaftsanalytische Methode) as “the only way that affords
us an understanding of the contribution of our reason to knowledge”.76
Its entire purpose is to draw a sharp distinction between the “subjective” role
in cognition of concepts and principles (“the contribution of reason”) from the
“objective” reality to which they are coordinated and hence “constitute” as an
object of knowledge. Significantly, Reichenbach holds that the “constitutive”
role of concepts and principles, what he calls “the contribution of reason to
knowledge” is a subjective one.77 With respect to geometry, relativity has shown
that freedom to make coordinate transformations (general covariance) reveals
the underlying objective character of the metric:

‘%e J. Ehlers, F.A.E. Pirani and A. Schild, ‘The Geometry of Free Fall and Light Propagation’,
in General Relativity; Papers in Honour of J.L. Synge, L. O’Raifeartaigh (ed.), (Oxford, Clarendon
Press, 1972) pp. 63-84.
‘sFor a critical discussion, see A. Griinbaum, ‘General Relativity, Geometrodynamics and
Ontology’, chapter 22 of Philosophical Problems of Space and Time, 2nd enlarged edition
(Dordrecht/Boston. D. Reidel. 1973). DD. 730ff. and also the 1983 paper of Stachel (op.cit., note 15).
A comprehensive assessment of Weyi’s chronogeometric method is given by J. Ehiers, ‘Hermann
Weyl’s Contributions to the General Theory of Relativity’, in W. Deppert el al. (op.&., note 49).
Ehlers notes (p. 88) that ‘the most obvious local properties of the system of light rays and freely
falling part& do not...give rise to a pseudo-Riemannian metric, they lead to a Weyl geometry.’
76Relativitiitstheorie und Erkenntnis A Priori (Berlin, J. Springer, 1920) p. 71; in the English
translation by Maria Reichenbach, The Theory of Relativity &d 2 Priori Knowledge (Berkeley and
Los Angeles, University of California Press, 1965), this passage is at p. 74.
“See my “ ‘P(ointtC(oincidence) Thinking’: The Ironical Attachment of Logical Empiricism to
General Relativity (And Some Lingering Consequences)“, Studies in the History and Philosophy of
Science 23 (1992) 471497.
858 Studies in History and Philosophy of Science

The theory of relativity teaches that the metric is subjective only insofar as it is
dependent upon the arbitrariness of the choice of coordinates, and that independently
of these it describes an objective property of the world.78

However, just two years later, Reichenbach reverses his assessment of the
epistemic polarities governing knowledge of the metric in GTR from
“objective” to “arbitrary”:

It is the significance of the theory of relativity to have discovered the limits of


arbitrariness. According to the theory of relativity, the choice of a geometry is
arbitrary; but it is no longer arbitrary once congruence has been defined by means of
rigid bodies.79

The metric is no longer “an objective property of the world”; instead,


“congruence (is to be) defined by means of rigid bodies.’ How is this to be
done?
Details are provided already in another, though little known, paper of
1922 through the introduction of the concept of “force of type x”, which
has two fundamental properties: it acts in the same manner on all bodies
regardless of their composition, and there are no insulating partitions
shielding bodies against it.80 Reichenbach then shows that, using the notion
of “force of type X”, an empirically equivalent theory (combining Euclidean
geometry and these unknown forces) can be developed which is a rival to the
non-Euclidean geometry of Einstein’s general relativity. It is, Reichenbach
acknowledges, “simpler” to set X = 0, a choice that can be considered to be
well-founded experimentally. Still the resulting definition of congruence is
nonetheless arbitrary. However, the concept of “force of type X” also
enables Reichenbach to define the concept of transport of a body “without
modification”:

One says that a body is transported without modification if:


1. It does not submit to known physical forces.

“The Theory of Relativity and A Priori Knowledge (opcit., note 76), p. 90.
79‘Der gegenwirtigen Stand der Relativitiitsdiskussion. Eine kritische Untersuchung’, Logos 10
(1922), 316-378, at p. 368; translated, with omission of “some minor criticisms of relativity which
are of no historical importance” as ‘The Present State of the Discussion on Relativity’, by Maria
Reichenbach, as reprinted in M. Reichenbach and R. S. Cohen (eds), Hans Reichenbach Selected
Writings, Volume 2 (Dordrecht, D. Reidel, 1979), pp. 347 at p. 38. Interestingly, the three pages
of discussion of Weyl’s theory in the original at pp. 365-368, are among those chosen to be excised
in the translation as “of no historical importance”. The assessment of Weyl’s theory here notably
contrasts with Reichenbach’s subsequent opinion: ‘Certainly Weyl manages the explanation of
the univocal transportability of natural measuring rods only very incompletely. However, that
Weyl has sought to take this path remains, independently of the empirical correctness of
his theory, an ingenious advance (genialer VorstoJ) in the philosophical foundation of physics.’
(p. 368.)
‘“‘La signification philosphique de la thCorie de la relativitk’, Revue Philosophique de la France et
de l’ktranger 94 (1922), 5-61; the paper was translated into French by “M. Leon Bloch”. The
characterization of a “force d’espkce x” is on p. 36.
Weyl, Reichenbach and the Epistemology of Geometry 859

2. One does not admit any forces of type X in the comparison of its length at
different places.81

With this definition of a rigid body as a body that can be transported without
modification, the arbitrariness of geometry is exhausted, and the question of
knowing which geometry results in the ensuing investigation of spa@-time)
acquires an empirical nature.** But this definition is not necessary and one may
just as well choose an arbitrary geometry while allowing that there are forces of
type X. “This is the road”, Reichenbach revealingly notes, “that Weyl has
travelled with a perfect rigor”.*3
The case made here for an attenuated conception of geometric empiricism is
largely identical to Reichenbach’s “mature” view of the epistemology of
geometry, as set out six years later in the classic Philosophie der Raum-Zeit
Lehre. What happened in the short space of two years to so fundmentally
change Reichenbach’s mind on the status of the metric in GTR?
To be sure, Albert0 Coffa first called attention to a letter from Schlick to
Reichenbach of 26 November 1920 in which Schlick points out that the
constitutive role played by Reichenbach’s “principles of coordination” is none
other than that played by conventions in PoincarC’s account of geometry;
subsequently Reichenbach will call his coordinative principles “coordinative
definitions”.*4 However, a more compelling motivation for Reichenbach’s
transformation from a relatively straightforward geometric empiricism in 1920
to a form of conventionalism in 1922 lies in the hypothesis that he is responding
to Einstein’s lecture “Geometry and Experience” of January 1921. For
this lecture is above all notable for the pragmatic defense given there of
what Einstein terms “practical geometry” against an alternative position
that upholds an essential inseparability of geometry and physics and which
Einstein, with some liberty, takes to be inherent in or implied by Poincark’s
conventionalism.
In actual fact, it would seem that the alternative which Einstein has in mind
is not Poincart’s, given the non sequitur-if PoincarC’s position is indeed the
holistic position of Einstein’s characterization (see immediately below)-that
“Ibid., p. 37: ‘On dira qu’un corps est transport6 sans modification si: 1. 11 ne subit aucune des
influences physiques connues; 2. On n’admet aucune force d’espi-ce X pour la comparaison de sa
grandeur en des lieux diffkrents.’
**Zbid. pp. 37-38: ‘La gkomttrie est arbitraire tant que l’on ne fait pas de convention sur le
concept ‘sans modification’ mais on peut prkiser ce concept d’une faGon purement physique en
relation avec les propriCtCs des forces et sans l’emploi de la gComCtrie. Le corps rigid peut Ctre dkfini
d’une faGon naturelle, sans qu’il soit ntcessaire pour cela de poser une gtomktrie normale. Une fois
ceci fait, la question de savoir quelle gkomktrie s’ensuit est de nature expkrimentale. Mais si l’on n’a
par fait cela, la gkomttrie, comme aussi la forme de tout corps physique, demeurent indCtermintes.’
8’Zbid., pp. 40-41: ‘Mais il nous faut encore une fois insister sur ce point: la definition qui a CtC
donnCe du corps rigide n’est pas n&xsaire. On peut aussi bien admettre des forces d’esptce X et
prescire une gtomktrie arbitraire....C’est le chemin qu’a suivi avec une rigueur parfaite Weyl.’
841n Linda Wessels (ed.) The Semantic Tradition from Kant to Carnap; To the Vienna Station
(Cambridge and New York: Cambridge University Press, 1991), pp. 201-202. This text was written
before the end of 1984.
860 Studies in History and Philosophy of Science

we should always choose Euclidean geometry as simpler, but Weyl’s, as is made


apparent in an ensuing discussion which rehearses, in this public forum,
Einstein’s “prehistory” (constancy of spectral lines) objection against Weyl’s
theory. “Practical geometry” is formulated thus:

Solid bodies are related, with respect to their possibilities of arrangement, as are
bodies in Euclidean geometry of three dimensions; then the propositions of Euclidean
geometry contain expressions concerning the behavior of practically rigid bodies.*j

and Einstein’s case for its validity rests upon the supposition of the existence
of “practically rigid” bodies, that is, bodies on which “two ‘tracts’ (‘Strecke’)
found to be equal once and anywhere are equal always and everywhere”. This
is just to assume that congruence relations are maintained under transport of
measuring rods (or clocks), an assumption for which, Einstein notes, “the
existence of sharp spectral lines comprises a compelling empirical proof”.86
Einstein goes on then to examine the criticism he takes to be PoincarC’s:

Why is the equivalence of the practically rigid body and the body of geometry-which
suggests itself so readily-denied by Poincarir ? Simply because under closer
inspection, the real solid bodies in nature are not rigid, because the geometrical
behavior...depends upon temperature, external forces, etc. Thus the original,
immediate relation between geometry and physical reality appears destroyed, and we
feel impelled to the following more general view, which characterizes PoincarC’s
standpoint.

Geometry (G) says nothing about the relations of real things, but only geometry
together with the system of physical laws (P). Using symbols, we may say that only
the sum of G + P is subject to the control of experience. Thus, G may be chosen
arbitrarily, and also parts of P; all these laws are conventions. All that is necessary to
avoid contradictions is to choose the remainder of P so that G and the whole of P are
together in accord with experience.... Sub specie aeterni PoincarC, in my opinion, is
right. The idea of the measuring rod and the idea of the clock co-ordinated to it in the
theory of relativity do not find their exact correspondence in the real world. It is also
clear that the rigid body and the clock do not play the part of irreducible elements
in the conceptual edifice of physics, but that of composite structures which should
not play any independent part in theoretical physics. But it is my conviction that
in the present stage of development of theorectical physics these concepts must still
be employed as independent concepts; for we are still far from possessing such
certain knowledge of theoretical principles as to be able to give exact theoretical
constructions of solid bodies and clocks.87

This lengthy passage makes several points relevant to our discussion. The
first is that because truly “rigid” bodies do not exist in nature, PoincarC is in
principle right: only geometry and physics combined yield testable statements

85”Geometrie und Erfahrung” (op.&., note 4), p. 125.


86Ihid.,
pp. 127 and 128: ‘Die Existenz scharfer Spektrallinien bildet einen iiberzeugenden
Erfahrungsbeweis fiir den gennanten Grundsatz der praktischen Geometrie.’
x71hid.,pp. 126127.
Weyl, Reichenbach and the Epistemology of Geometry 861

regarding the metric of space-time. 88 The second is that at the present stage of
the development of physics, the notions of rigid rod and clock are adopted as
convenient tools of a “practical geometry”, although this is only to be viewed as
a temporary expedient.a9
However, squaring the Poincare of Einstein’s discussion with positions
identifiably held by Poincart has long been recognized as a problem of
interpretation; and several commentators have suggested that it is not so much
Poincart’s conventionalist critique of “geometric empiricism” but rather
Duhem’s holist epistemological perspective that Einstein holds to be correct sub
specie aterni.90 But as suggested above, the scientific context of Einstein’s
defense of “practical geometry” is that of the challenge posed by Weyl’s
non-Riemannian alternative to the congruence erngeometrische) assumptions
of the pseudo-Riemannian metric of GTR.
The inseparability of geometry and physics in this context is not then a view
held on the grounds of some general epistemological position (“holism”);
rather, it is the inevitable outcome of the attempt to find a unified account of all
physical phenomena fully in terms of the dynamical interaction of matter and
the metric of space-time, the only perspective from which GTR can be seen as a
completed theory. With a dazzling display of mathematical brilliance, Weyl had
(prematurely, it seems, in Einstein’s opinion) forced the issue of the necessary
“expansion” of GTR, but we must not overlook that Einstein did not have to
be convinced of the eventual necessity of such a move.91 For Reichenbach, on
the other hand, this attempt to seemingly “geometrize” physics precisely
threatens the methodological analysis of science he has proposed by collapsing
**In 1922, commenting upon Einstein’s discussion here, Reichenbach observes: ‘It is not possible
to define solid body (feste K&per) without referring to physical laws. These laws can also be chosen
in such a way that the rigid body (starre K&per) is de$ned through Euclidean geometry...But in that
case a restricting condition enters into the choice of physical laws. Hence, according to Einstein’s
formulation, only the sum G + P of geometry and physics is susceptible to the control of
experience.’ ‘Der gegenwartige Stand der Relativitatsdiskussion. Eine kritischen Untersuchung’
(op.cit., note 79), pp. 354355.
“Einstein returns to the subject of this “necessary evil” in his ‘Autobiographical Notes’ in P. A.
Schilpp (ed.), Albert Einsrein: Philosopher-Scientist (Evanston and Chicago, Northwestern Uni-
versity Press, 1949) pp. 2-95, at pp. S-61, pungently observing: ‘...it was better to permit such
inconsistency-with the obligation, however, of eliminating it at a later stage of the theory. But one
must not legalize the above-mentioned sin so far as to imagine that intervals are physical entities
of a special type, intrinsically different from other physical variables.’
90A. Griinbaum, Philosophical Problems of Space and Time (op.cit., note 75), chapter 4 (‘Critique of
Einstein’s Philosophy of Geometry’), and D. Howard, ‘Einstein and Duhem’, Synthese 83 (1990),
363-384. Of course, it is the holism of Einstein’s discussion, rather than any views that Duhem might
have had on the epistemology of geometry, that prompts these interpretative efforts, To be sure,
Grtinbaum (p. 135) also notes Weyl’s holism, but as being “an endorsement of Duhem’s position”.
“Recall here Einstein’s remarks in his ‘Autobiographical Notes’: ‘It seemed hopeless to me at
that time [i.e., of the first formulation of GTR-TR] to venture the attempt of representing the total
field [“general field (in which quantities corresponding somehow to the electromagnetic field occur,
too)“] and to ascertain field-laws for it.” And “If one had the field-equation of the total field, one
would be compelled to demand that the particles themselves would everywhere be describable as
singularity-free solutions of the completed field-equations. Only then would the general theory of
relativity be a complete theory.” (op.cit., note 89) pp. 73 and 81.
862 Studies in History and Philosophy of Science

its sharp distinction between the subjective or arbitrary component of our


knowledge and the objective, empirically determined, remainder, a distinction
he will view as “the philosophical significance” of the theory of relativity.
In his Philosophy of Space and Time, first published in 1928,92 Reichenbach
gives a highly influential argument, the intent of which is to provide a rigorous
definition of the notion of a “rigid body”, thus undergirding, in effect, the
“practically rigid body” of Einstein’s discussion by stipulating an equivalence
between certain physical bodies (measuring rods and clocks) and chrono-
geometric concepts. Unlike the general tenor of Einstein’s essay, however, the
aim here is to unambiguously separate the whole that is geometry and physics
through use of “coordinative definitions”, the conventional equivalences just
‘*Philosophic der Raum-Zeit-Lehre (Berlin, DeGruyter, 1928) especially Sections 3-8; translated
by M. Reichenbach and J. Freund, with omission of an appendix on Weyl’s theory, as The
Philosophy of Space and Time (New York, Dover, 1958). This edition was published with an
‘Introduction to the English Edition’ by Rudolf Carnap who writes (p. vi): ‘This work was an
important landmark in the development of the empiricist conception of geometry. In my judgement
it is still the best book on the subject.’ The original German edition contained a 42 pp. Anhang on
Weyl’s theory, not translated for the English edition, which both confirms the significance of Weyl’s
theory in the formulation of Reichenbach’s epistemology of geometry and at the same time reveals
a rather unmodulated and dogmatic streak in the latter. Reichenbach’s ostensible intention is to
show that the “geometric significance of electricity”-unlike that of gravitation-has led to nothing
of independent physical interest and that therefore Weyl’s ambition to do physics by finding the
“most natural geometry” is an, as yet, unwarranted and arbitrary attitude. To show this he
concocts his own “unified theory“, essentially following Eddington in initially assuming a
“fundamental tensor” G,. = gPV+&. whose two parts are a symmetric part representing the
gravitational field and an antisymmetric part representing the electrical field. The metric of the
space-time manifold is then defined in a manner that omits the antisymmetric part:
d? = G,,dx,dx, = g,,.dx,dx,. The metric is physically “realized” by measuring rods, clocks and
light rays (these are its “indicators” to which metrical concepts are “coordinated”). Next, a “shift
operation”, producing a “shift field” (Verschiebungsfeld), fiv is defined in terms of the gflVand the
h,, that is, solely in terms of the G,“. Finally, a law of motion for an arbitrarily charged particle of
unit mass is defined in terms of the I;,,, which is shown to be equivalent to its counterpart in GTR.
This enables Reichenbach to claim that the “shift” is “realized” by such a particle, with the
consequence that “straightest” lines, physically realized by such particles, will be different from
“shortest” lines, realized through the measuring rods coordinated to the metric. Only if the charge
is 0 on a unit mass point will “straightest” lines be at the same time “shortest” lines. Though as far
as I know Weyl never responded to this argument of Reichenbach’s (and by 1928 when
Reichenbach’s book was published, Weyl had given up arguing for his theory, explicitly rejecting
it in 1929-see note 104 below), there are two overriding reasons why he would not have been
persuaded. First, Reichenbach’s insistence that the “physical realizations” of the g,,, and I;” fields
be established by “coordinative definitions” violates the espirit of Weyl’s “systematic” alternative
to GTR (see the concluding remarks of this section). Second, and most importantly, the objections
he raised in 1923 against Eddington’s theory apply straightforwardly to Reichenbach’s construc-
tion: ‘Any theory, which deduces the gPV out of other, more fundamental state magnitudes, has a
great objection against it: how should rt explain that the quadratic form formed from the gwVnever
degenerates and generally possesses the same inertial index [i.e., one negative sign-TR]? In our
conception, group-theoretic investigation shows that this requirement lies in the essence ( Wesen)
of the metric.... If one robs the metric of its fundamental (urspriing[ichen) character, then in my
opinion experience gives no basis any more for ascribing an affine connection to the world.” Here
Weyl alludes to the Grundtatsache of “pure infinitesimal geometry”, that the metric univocally
determines the affine connection and hence that “shortest” and “straightest” geodesics must
coincide. For these reasons, Weyl concluded, the Eddington theory was simply “undiscussible”
(undiskuteirbar) (Raum-Zeit- Materie, 5 Edn, p. 324). Interestingly, Reichenbach appears to have
consulted only the 3rd (1919) edition of this work, in which there is, of course, no mention of
Eddington’s theory.
Weyl, Reichenbach and the Epistemology of Geometry 863

mentioned. These are held to “exhaust” the “arbitrariness” involved in


determining the geometry of space-time, hence, rendering this question empiri-
cally determinate. In order to define a rigid body, Reichenbach appeals to a
distinction between what he terms “differential” and “universal” forces.
Dz@rentiul forces are those forces that we know are responsible for altering
or deforming a measurement body, e.g. heat. We know that an aluminum meter
stick is slightly longer on a very warm day than on a very cold one. And we can,
and do, correct for these fluctuations. For example, if we know the coefficient
of thermal expansion of aluminum, we can say how much longer the stick is at
98 F than at 0 F. And we know that the same temperature change will alter an
aluminum meter stick more than one made of oak. We also know that we can
insulate measuring bodies against such deforming forces. Universal forces, on
the other hand, affect all bodies in the same way and cannot be screened off with
insulating walls or other devices; though their earlier debut here remains
unacknowledged, these are just the “forces of type X” of 1922.
How could we ever detect whether a measuring body was under the influence
of universal forces? The short answer, according to Reichenbach, is that this is
“logically impossible” to know. 93 Suppose two measuring rods are determined
to be congruent in one region of space-time (A) and then one is transported to
another region (B) and, after correcting for any suspected influence of
differential deforming forces, we ask if it is still congruent to the first rod which
remained behind at A. Is the length measured at B congruent to “the same”
length measured back at A? We cannot know. And the fact, if it is a fact, that
we cannot know does not mean that it is impossible that the length of the rod
altered during transport. Thus we are faced with the following situation which
is superJciaZZy similar to that of Poincare in Einstein’s characterization.94

“Ibid. Section 7 (‘Technische Unmoglichkeit und prinzipielle Unmiiglichkeit’).


94The problem of undetectable “deforming” forces afflicting the characterization of the geometry
in Poincare’s “thermal world” Gedankenexperiment (op. cit., note 24) at pp. 65-68, has been often
taken as the immediate inspiration for Reichenbach’s distinction between “differential” and
“universal” forces; see, e.g. L. Sklar, Space, Time, and Spacetime (Berkeley and Los Angeles,
University of California Press, 1977), pp. 88-101. However, there are several striking disanalogies
in Reichenbach’s discussion: (1) Poincare’s fanciful example is expressly counterfactual (all bodies
have the same coefficient of thermal expansion and are in instantaneous thermal equilibrium with
their surroundings); whereas in Reichenbach, there is 110fact of the matter; only “descriptive
simplicity” (see below) is involved in choosing or not to appeal to universal forces; (2) the
“deforming” forces in Poincart are thermal but undetectable (since they affect everything in
the thermal world); in Reichenbach, thermal forces are “differential” (and hence correctable);
(3) Reichenbach intends his distinction to conventionally distinguish what “belongs in physics” and
what “belongs in geometry” (Philosophy of Space and Time, op.cit., note 92, p. 27) whereas the
whole point of PoincarC’s exercise is to deny that a principled distinction can be made here: we
simply choose the “simplest” geometry. While Reichenbach’s distinction may have been implicit in
Poincart’s thermal world Gedunkenexperimenf, his discussion nonetheless pointed to a different
conclusion, as Reichenbach, in one of his few references to Poincart, explicitly notes: “He (i.e.
Poincare) overlooks the possibility of making objective statements about real space in spite of the
relativity of geometry,...” (Philosophy of Space and Time p. 36, footnote 3). In my view, the concept
of “universal forces” makes sense only as employed as a counter to Weyl’s theory.
864 Studies in History and Philosophy of Science

Only the combination of geometry (G) and universal forces (F) yields a
testable statement regarding the congruence relation between the two rods at A
and B. There are two possibilities entertained by Reichenbach. The first is that
there are universal forces (F does not equal 0); hence we can choose any
arbitrary geometry we like and explain a resulting deviation from G (e.g. a
supposed alteration of length at B) as due to the effects of universal forces. The
second is to define a “rigid body” as a body on which only differential, i.e.
detectable and correctable, forces act. There are no universal forces. Hence our
geometry G alone gives a univocal determination of spatiotemporal reality,
and we can say that the rods at A and B are congruent.
We are free to choose, Reichenbach coyly informs us, either combination, a
choice which he terms “the relativity of geometry”. For example, in order to
“save” any particular favored geometry, we need only postulate the existence of
non-zero universal forces. Of course, given the shadowy standing of universal
forces,95 it seems certain that anyone in his right mind will opt for the latter.
Because it is “logically impossible” to empirically determine whether rods and
clocks are altered under transport, we are “naturally” led to adopt a convention
that they are not altered. This choice is tantamount to the standard definition of
congruence and (perhaps borrowing Eddington’s term) it singles out a “natural
geometry” from a “class of equivalent descriptions”.96 Proceeding thus, an
“objective statement” about the “geometry of real space” is rendered as “a
statement concerning the relation between the universe and rigid measuring
rods”.97 Still, non-standard geometries are equally compatible with the
observed behavior of the measuring instruments; selection of one of these (with
the accompanying appeal to universal forces) merely amounts to changing one’s
coordinative definition of “same length” for bodies at a finite distance apart.
Only “descriptive simplicity”, which “has nothing to do with truth”, is involved
in such a choice.98 Not any particular choice, but that a choice be made, is
necessary. This is not an idle exercise. Indeed, Reichenbach has it that

“Reichenbach’s distinction between “differential” and “universal” forces has puzzled, among
others, Griinbaum, Philosophical Problems of Space and Time (op. cit., note 75) who (pp. 92-94)
concludes that the term “universal forces” is only a misleading metaphor, and R. Torretti,
Relativity and Geometry (Oxford, Pergamon Press, 1983) p. 238, who labels them as “science
fictions” having no place in “real science”: “The truth is that a ‘universal force’, a force which
cannot be detected by any means because it modifies the shape of the instruments of detection in
the exact way required to conceal its presence, belongs to the realm of science fiction and cannot
be seriously countenanced in real science. At least in physics, a difference to be a difference must
make a difference.”
96The term “natural geometry” is used in The Rise c~fScientificPhilosophy (Berkeley and Los
Angeles, University of California Press, 1951) p. 134. while that of “class of equivalent
descriptions” appears in Philosophical Foundations of Quantum Mechanics (Berkeley and Los
Angeles, University of California Press, 1944) p. 19.
97Philosophie der Raum-Zeit-Lehre (op.cit., note 92), p. 50: ‘...gibt es eine objektive Aussage iiber
die Geometrie des wirklichen Raumes: sic ist eine Aussage iiber eine Beziehung zwischen dem
Universum und starren Maj3stiiben.’ (original emphasis].
981bid., p. 47 (English translation, p. 35).
Weyl, Reichenbach and the Epistemology of Geometry 865

One can characterize the philosophical achievement (Leistung) of the theory


of relativity in this way, that it has demonstrated the necessity for metrical coordin-
ative definitions in many places in which one had previously sought cognitions
(Erkenntnisse).

Pointing forward to the metaphilosophical lesson of the method of “equiva-


lent descriptions”, Reichenbach warns that “pseudo-problems arise if we look
for cognitions where definitions belong”, and he singles out Helmholtz for
praise in creating the “philosophical foundations” of the problem of geometry
by having “recognized the connection of the problem of geometry with that of
rigid bodies”.99
There has long been a puzzling, and seemingly incongruous, element in this
argument’s appeal to a shadowy distinction between “universal” and
“differential” forces in order to establish the necessity of making coordinative
definitions, which thus render cognition empirically determinate. The distinc-
tion, Reichenbach tells us, is just that between what belongs to geometry
(“universal”) and to physics (“differential”). But then in a passage occurring
much later in his book, 100Reichenbach identifies gravity as a “universal force”
since (as is postulated by the Principle of Equivalence), “It does indeed affect all
bodies in the same manner.“101 Commentators have been severely critical of
Reichenbach on this identification, claiming, for instance, that Reichenbach has
not succeeded in establishing a non-metaphorical sense for the term “universal
force”, a failing which appears to lead him into a self-contradictory position.lo2
What must be remarked, however, is that the context of this passage is a
discussion of the philosophical significance of the identification of gravitation
and metric in Einstein’s field equations, the problem thus being posed of
whether this identification “deprives gravitation of its physical character”.
Though no representatives of this view are implicated by name in this text,
Reichenbach’s animus here, in both content and in language, is clearly against
Weyl’s claims of the “geometrization of physics” that we encountered in
Section 2 above.
However, the “relativity of geometry” Reichenbach defends in the initial
sections of his book cannot provide him with a knock-down argument against

991bid., pp. 24 and 48 (English translation, differing from the above, pp. 15 and 36). Reichenbach
repeats this assessment of the philosophical significance of the theory of relativity again in a work
published the next year: ‘the theory of relativity..., whose epistemological significance precisely
consists in having revealed as subjective elements earlier held to be objective’. ‘Ziele und Wege der
physikalischen Erkenntnis’, chapter 1 (pp. l-80) of Band IV (Allgemeine Grundlagen der Physik), of
H. Geiger and K. Schell (eds), Handbuch der Physik (Berlin, J. Springer, 1929) p.44. An English
translation of this article by Peter Heath appears in Hans Reichenbach Selected Works, Volume
Two (opcit., note 79) pp. 120-225; the cited passage is slightly differently translated at p. 174.
‘OOPhilosophie der Raum-Zeit-Lehre, pp. 294296; Philosophy of Space and Time, pp. 256258.
“‘Of course, it is precisely gravitational “forces”, in the guise of the scalar of world curvature,
which have been posited by Weyl as the field strengths to which the electrical forces within the
constituent atoms of measurement instruments “adjust”.
“*See Griinbaum (op.cit., note 75) p. 93.
866 Studies in History and Philosophy of Science

those who claim a “geometrization” of gravity. As a result, to argue against


such a “geometrization” he essentially suspends the conventionality of congru-
ence definitions as soon as the discussion turns from shadowy universal forces
to the physically real gravitational field. For, on the one hand, he now argues,
“the universal effect of gravitation upon all kinds of measuring instruments
defines a single geometry”, though this is not allowed to mean that “the
measuring instruments are changed by the gravitational field” but only that
“the instruments are ‘free from deforming forces’ in spite of gravitational
effects”. This is to conveniently overlook the conventional character of the
definition of “rigid body” and hence not to permit a non-standard definition of
congruence, supplemented by appeal to deforming “universal forces”. On the
other hand, Reichenbach goes on to maintain that it is nonetheless necessary
“to invoke a force as a cause” for the fact of the observed “general
correspondence of all measuring instruments”. In this sense, “the gravitational
field has the physical reality of a field of force”. But crucially, the force field
is to be regarded as “the cause of geometry itself, not as the cause of
the disturbance of geometrical relations”. Again, though the “relativity of
geometry” would appear to allow the latter to be just as viable a choice as the
former, only the standard congruence definition is permitted.
What becomes clear in reflecting on these passages is that, for the crucial case
of the gravitational fieldeven as it is recognized as a “universal force”-
Reichenbach essentially revokes the latitude offered by the “relativity of
geometry”. His position against an alleged “geometrization” of physics is
attained by the merest fiat: “Geometric measurement is a handling of indicators;
therefore, the metric of the indicators is at once the measure of the field that
determines their adjustment (Einstellung). “103
But now recall, from Weyl’s perspective, the consequences of adopting the
“purely infinitesimal” conception of geometry and the ensuing notion of a
more general (Weyl) metric. This has, as yet, “nothing to do” with the
behavior of instruments of measurement. The problem is then to accommodate
the fact that we do observe invariance of congruence under transport of
measuring rods and clocks. But this is not to be viewed as a primitive fact that
does not warrant, or is incapable of, explanation. Here is an apparent
transgression of that application of the principle of sufficient reason which
motivates verificationism and Reichenbachian conventionalism: the principle
tells us that if rods and clocks are observed to be unaltered under transport,
then we infer, or rather stipulate, that indeed they have been unaltered: there is
nothing further to be explained or accounted for. This, of course, appears to be
a reasonable inference supported by masses of ordinary experience. But, in
principle, it is not an option for one attempting to derive the behavior of
““Philosophic der Raum-.&ii-Lehre, p. 296; Philosophy of Space and Time, p. 258; emphasis in
original.
Weyl, Reichenbach and the Epistemology of Geometry 867

measuring instruments from a “systematic theory” of matter-field interactions.


In this regard, Reichenbach’s methodological imperative to legislate
a sharp “subjective”/“objective”, convention/fact dichotomy in the analysis
of scientific theories miscasts Weyl’s proposal for a “systematic” conception of
GTR as merely an unnatural, but alternative “choice” from the class of
equivalent descriptions regarding the geometry of space-time.
By the same token, Reichenbach’s “equivalent descriptions” characteriz-
ation fails to do justice to the pragmatically provisional nature of Einstein’s
defense of “practical geometry” in “Geometry and Experience” and it entirely
effaces the dissatisfaction, expressed there, of this manner of conventionalist
solution to the problem of the co-determination of metric and field. “Equiva-
lence of descriptions” is purchased at the prohibitively high cost of a gross
comparison of space-time theories on the meagre ground provided by the
imperfect indications of measurement instruments. But this kind of verifi-
cationism essentially ignores that GTR fundamentally transforms the very
notion of a “measurement instrument” in as much as such an object should
have, in principle, no independent existence apart from the reality of the physical
fields into which it is “plunged”. Moreover, the application of verificationism
here washes out all of the details of interest in this instance of theory
comparison, as is testified by the total absence of any trace of this comparison
in the literature on Reichenbach’s views on the epistemology of geometry.
Finally, the invocation of “universal forces”, of course, serves not just the
epistemological purpose of setting the table for a neo-conventionalist
“solution” to the problem of “the epistemology of geometry”. It also served at
the time a clearly rhetorical purpose in discrediting Weyl’s approach as, in an
important sense, being unphysical, indeed, near to the point of ridicule. This
dialectical context of Reichenbach’s geometric conventionalism has been
forgotten, but its philosophical significance today should be clear, not least for
challenging the loaded deck of a rigid separation of convention and fact which
logical empiricism claimed as the epistemological legacy of the theory of
relativity.

5.
What do we learn from this episode? And what, ultimately, can we therefore
say of the problematic of “the epistemology of geometry”? What kind of
question is it whose traditional spectrum of possible answers (a priorism,
conventionalism, empiricism), to one extent or another, all disguise the full
complexity of the intertwining of geometrical and physical questions in a theory
such as GTR? Indeed, all presuppose that it makes sense to inquire of
“geometry” in the sense of “geometry of physical space”, although GTR
teaches that this is not a well-defined question since it cannot be exclusively
answered in its own, i.e. geometrical, terms.
868 Studies in History and Philosophy of Science

Weyl and Einstein, following on the path begun by Riemann saw most deeply
into the unity that is geometry and physics and into the artificiality of
attempting to distill the one from the other. This is the philosophical lesson of
GTR, although now GTR is seen, as it was by Einstein and Weyl, as requiring
completion through an extension to encompass a theory of matter. As it
happened, neither Weyl’s nor Einstein’s approach to unification survived the
quantum theory, although only Weyl would concede this.‘04
And there is a poignant irony in that Weyl, who was, even more than Hilbert,
perhaps the stellar exponent of the pure mathematician’s approach to physical
theory, gradually came to become sceptical of this method, even as Einstein
took it up. Voicing in private his “unofficial” criticism around 1919 or 1920,
Einstein had remarked to Weyl: “Well, Weyl, let’s leave it alone. For in such a
speculative way, without a guiding intuitively physical principle, one cannot do
physics!“‘05 Toward the end of his life, sobered by the profusion of elementary
particles revealed in high-energy physics, Weyl was ready to agree: physics was
not yet ready for such a deeply speculative approach to unification.106 But
Einstein, having tasted the forbidden fruit of pure mathematics, was never to
look back, spending years in the wilderness in pursuit of a unified theory of
fields.

“““By this new situation, which introduces an atomic radius into the field equations
themselves-but not until this step-my principle of gauge invariance, with which 1 had hoped to
relate gravitation and electricitry, is robbed of its support.’ ‘Gravitation and the Electron’, Rice
Institute Pamphlets XVI (October, 1929) pp. 280-295; at p. 284. By “atomic radius” Weyl intends
the length hlmc (“the wave length of the electron”), which is seen as “an absolute length” and he
goes on in the next sentence to acknowledge that his gauge principle continues to play a role in the
new quantum mechanics as a phase factor of the i+~function (the “matter field”), completing the
analogy that had been shown by F. London, ‘Quantenmechanische Deutung von der Theorie von
Weyl’, Zeitschrift fir Physik 42 (1927). 375-389 (for discussion, see V.V. Raman and P. Forman
‘Why was it Schrodinger Who Developed de Broglie’s Ideas? Historical Studies in the Physical
Sciences 1 (1969) 291-314, esp. pp. 303ff.). In this quantum mechanical context, the gauge
transformations are:
v+e”y/, where v is the “potential of the matter field” and p,+v, - hle~%.l&J, where v, the
electromagnetic potential (=CA,dx-” in contemporary notation) and e the charge of an electron
citing here the version from Weyl’s The Theory of Groups and Quantum Mechanics, translated from
the second (1931) German edition by H. P. Robertson (New York, Dover, 1950; first published by
Methuen and Co, 1931) p. 100. The principle of gauge invariance is now seen as linking not
gravitation and electricity, but electricity and matter. There is an ironic twist to this development.
As C. N. Yang has recently pointed out (‘Hermann Weyl’s Contribution to Physics’, in K.
Chandrasekaran (ed.), Hermann Weyl, 1885-1985; Centenary Lectures at the ETH Ziirich (Berlin
and New York, Springer, 1986) pp. 7-21), viewing Weyl’s gauge transformation for a, now as a
phase transformation into Weyl’s original gauge transformation, renders Einstein’s “prehistory”
objection nugatory as it now concerns a phase difference which does not influence, e.g. the rates of
two clocks transported along different paths. Not only does Einstein’s objection disappear, but such
a phase difference (conceived as an interference experiment with two electrons) can actually be
measured, as was shown in an analogous situation by the experimental verification of the
Aharonov-Bohm effect by Chambers in 1960.
““‘Na 9 Weyl 7 lassen wir das! So-das heiDt auf so spekulative Weise, ohne ein leitendes,
anschauliches physikalisches Prinzipmacht man keine Physik!” Recalled by Weyl in a letter dated
19 May 1952 to Carl Seelig, as cited in Seelig’s Albert Einstein (Zurich, Europa, 1960) p. 274.
‘Oh‘Erkenntris und Besinnung’, Studia Philosophie XV (1955) 153-171.
Weyl, Reichenbach and the Epistemology of Geometry 869

And Reichenbach? With his antecedent insistence on epistemic foundations


that require a rigid separation of subjectivity and objectivity, convention and
fact, he was in no position to notice that the thin metaphilosophical gruel of
theoretical underdetermination stemming from conventionally generated
“equivalent descriptions” is no substitute at all, even qua “rational
reconstruction”, for the immensely rich tapestry against which this episode of
rival theories in science unfolded, and in which the philosophical significance
of general relativity stands forth.

Acknowledgements-I’d like to thank audiences at Northwestern University and the University of


Notre Dame, in particular, John McCumber at the former and James T. Cushing at the latter,
where earlier versions of this paper were read. Special thanks go to Arthur Fine, Don Howard,
David Joravsky, J. B. Kennedy, David B. Malament and two anonymous referees for valuable
comments and discussions on a previous draft.

Appendix: Summary Description of “Weyl Geometry”


This concise, modern sketch of a proof of Weyl’s result concerning the compatibility of the
metrical and affine structures of a differentiable manifold (“die Grundtatsache der infinitesimal
Geometric”-see Section 2 above) has been suggested by David B. Malament of the University of
Chicago. Notation follows Robert Wald, General Relativify (Chicago, University of Chicago Press,
1984).
We give two equivalent characterizations of “Weyl geometry”, and sketch how one gets from one
to the other. In what follows let M be a smooth, differentiable manifold.

First approach. One starts with a “Weyl structure” on M and proves the existence of a unique
compatible derivative operator on M.
Let w be a set of ordered pairs (gab p,) where, in each case, g,, is a smooth (non-degenerate)
semi-Riemannian metric on M, and q0 is a smooth, covariant vector field on M. Let us say that w
is a Weyl structure on M if the following two conditions are met:

(Wl) If (gob, p,) and (g’ab, q’,) both belong tow, there is smooth, everywhere positive, scalar field
R:M+IWsuch that
(i) g’,, = Qg,, and
(ii) I’, = q, - (l/R)(dD), [Here d is the exterior derivative operator on M]

(W2) If (gob, q,) belongs to w, and if Q:M+[W is a smooth, everywhere positive, scalar field on
M that satisfies (i) and (ii) in (Wl), then (g’ah. p’,) belongs to w.

Let w be a Weyl structure on M and let V, be a derivative operator (i.e. affine connection) on M.
We say that w and Va are compatible if for all pairs (gob, c,) in w,

It is easy to check that one pair in w satisfies (Cl) iff all pairs do. It is also easy to check that (Cl)
is equivalent to the following condition:
(C2) Given any smooth curve in M with tangent field q“, and any smooth vector field Y on the
curve that is constant (with respect to V,)

i.e. the rate of change of the (squared) length of y along the curve is proportional to the
(squared) length of 5” with proportionality factors ~~7~.
870 Studies in History and Philosophy of Science

Weyl’s theorem can now be cast as follows: given a Weyl stucture w on M, there is exactly one
derivative operator on A4 that is compatible with w.

Second approach. One starts with a conformal structure C on A4 and a derivative operator 8, on
M that is compatible with C (in a new sense of “compatible”); then one proves that there is exactly
one extension of C to a Weyl structure that is compatible with Vy (in the previous sense of
“compatible”).
Let us take a conformal structure on M to be a maximal set C of conformally equivalent, smooth,
(non-degenerate) semi-Riemannian metrics on M. Given a derivative operator Va on A4, we say
that C is compatible with V, if for all g,, in C, the following condition holds:

(C’l) There is a (unique) smooth covariant vector field q’, on A4 such that

It is easy to check that one metric in C satisfies (C’l) iff all metrics in C do. This condition admits
a second, more intuitive, formulation:

(C’2) (Length ratios with respect to g,, are preserved under parallel transport with respect to
V,.) Given any smooth curve in M with tangent field 7”. and any smooth vector fields y, ,ua on
the curve that are constant (with respect to V,), if g&&r is non-zero at one point on the curve,
then it is everywhere non-zero, and

The proof that (C’l) implies (C’2) is trivial. The converse requires just a bit of work; one shows
that if (C’Z) holds, and if n is the dimension of M, then

I, = (llnP’~,g,,

satisfies the condition in (C’l).

Now let C be a conformal structure on M compatible with VO. If we mate each metric g,, in C with
the field q,, identified by (C’l), then the resultant set of ordered pairs (gab, p,) is easily seen to be
a Weyl structure on A4 compatible with V0 (and, moreover, the only extension of C to a Weyl
structure on M that is compatible with V,).

Das könnte Ihnen auch gefallen