Quantum Theory and Functional Analysis
Quantum Theory and Functional Analysis
Quantum Theory and Functional Analysis
Klaas Landsman
Institute for Mathematics, Astrophysics, and Particle Physics (IMAPP),
Faculty of Science, Radboud University, Nijmegen, The Netherlands
and
Dutch Institute for Emergent Phenomena (DIEP), www.d-iep.org.
Email: landsman@math.ru.nl
arXiv:1911.06630v1 [math-ph] 15 Nov 2019
Abstract
Quantum theory and functional analysis were created and put into essentially their
final form during similar periods ending around 1930. Each was also a key outcome
of the major revolutions that both physics and mathematics as a whole underwent at
the time. This paper studies their interaction in this light, emphasizing the leading
roles played by Hilbert in preparing the ground and by von Neumann in bringing them
together during the crucial year of 1927, when he gave the modern, abstract definition
of a Hilbert space and applied this concept to quantum mechanics (consolidated in
his famous monograph from 1932). Subsequently, I give a very brief overview of three
areas of functional analysis that have had fruitful interactions with quantum theory
since 1932, namely unbounded operators, operator algebras, and distributions. The
paper closes with some musings about the role of functional analysis in actual physics.
Contents
1 Introduction 2
Epilogue 11
References 12
∗
To appear in the Oxford Handbook of the History of Interpretations and Foundations of Quantum
Mechanics, ed. O. Freire (Oxford University Press, 2021). This chapter suffers from a strict word limit,
as a consequence of which the discussion is often terse. For example, instead of explaining the technical
details, for which I refer to books like Landsman (2017), I have tried to sketch the relevant history at an
almost sociological level. I am deeply indebted to Michel Janssen and Miklos Rédei for helpful comments.
1 Introduction
Dijksterhuis (1961) concludes his masterpiece The Mechanization of the World Picture
(which ends with Newton) with the statement that the process described in the title
consisted of the mathematization of the natural sciences, adding that this process had
been completed by twentieth-century physics. As such, the topic of this chapter seems a
perfect illustration of Dijksterhuis’s claim, perhaps even the most perfect illustration.1
However, there is an important difference between the application of calculus to clas-
sical mechanics and the application of functional analysis to quantum mechanics: Newton
invented calculus in the context of classical mechanics,2 whereas functional analysis was
certainly not created with quantum theory in mind. In fact, the interaction between the
two fields only started in 1927, when quantum mechanics was almost finished at least from
a physical point of view, and also functional analysis had most of its history behind it.3
Functional analysis did have its roots in classical physics. Monna (1973), Dieudonné
(1981), and Siegmund-Schultze (2003) trace functional analysis back to various sources:
3. Integral equations, first studied by Abel in the 1820s and independently by Liouville
in the 1830s in connection with problems in mechanics. From the 1860s onwards
integral equations were used by Beer, Neumann, and others as a tool in the study
of harmonic functions and the closely related Dirichlet problem (which asks for a
function satisfying Laplace’s equation on a domain with prescribed boundary value,
and may also be seen as a variational problem). This problem, in turn, came from
the study of vibrating membranes via PDEs from the 18th century onwards.5
1
Einstein’s theory of General Relativity is an equally deep and significant example of the process
in question, but I would suggest that his application of Riemannian geometry to physics was, though
unquestionably an all-time highlight of science, less unexpected than the application of functional analysis
to quantum theory. Indeed, Riemann certainly thought about field theory and gravity in this connection.
2
The fact that Newton subsequently erased his own calculus from the Principia does not change this.
3
See Bernkopf (1966), Monna (1973), Steen (1973), Dieudonné (1981), Birkhoff and Kreyszig (1984),
Pier (2001), and Siegmund-Schultze (1982, 2003) for the historical development of functional analysis.
According to most authors this history occupied a period of about 50 years, starting in the 1880s and
ending in 1932 with the books by Banach, Stone, and von Neumann published in that year (see below).
4
See Monna (1973), Dorier (1995) and Moore (1995) for the history of linear structures.
5
It is hardly a coincidence that the Dirichlet problem was eventually solved rigorously in 1901 by none
other than Hilbert (Monna, 1975), whose role in functional analysis is described below and in §3 .
2
On the other hand, functional analysis benefited from–and was eventually even one of
the highlights of–the abstract or “modernist” turn that mathematics took in the 19th
century.6 In my view (supported by what follows), it was exactly this turn that made the
completely unexpected application of functional analysis to quantum theory possible, and
hence it seems no accident that Hilbert was a crucial player both in the decisive phase of
the modernist turn and in the said application. Thus Hilbert played a double role in this:
1. Through his general views on mathematics (which of course he instilled in his pupils
such as Weyl and von Neumann) and the ensuing scientific atmosphere he had cre-
ated in Göttingen.7 Hilbert’s views branched off in two closely related directions:
These came together in his Sixth Problem (from the famous list of 23 in 1900):8
2. Through his contributions to functional analysis, of which he was one of the founders:
By the depth and novelty of its ideas, [Hilbert (1906)] is a turning point in the
history of Functional Analysis, and indeed deserves to be considered the very first
paper published in that discipline. (Dieudonné, 1981, p. 110).
From both an intellectual and an institutional point of view, the connection between
quantum theory and functional analysis could be made (at least so quickly) because in
the 1920s Göttingen did not only have the best mathematical institute in the world (with
a tradition going back to Gauß, Riemann, and now Hilbert), but, due to the presence of
Born, Heisenberg, and Jordan, and others, was also one of the main centers in the creation
of quantum mechanics in the crucial years 1925–1927. It was this combination that enabled
the decisive contributions of von Neumann (who spent 1926–1927 in Göttingen, see §4).
It cannot be overemphasized how remarkable the link between quantum theory and
functional analysis is. The former is the physical theory of the atomic world that was
developed between 1900–1930, written down for the first time in systematic form in Dirac’s
celebrated textbook The Principles of Quantum Mechanics from 1930 (Jammer, 1989).
The latter is a mathematical theory of infinite-dimensional vector spaces (and linear maps
between these), endowed with some notion of convergence (i.e. a topology), either in
abstract form or in concrete examples where the “points” of the space are often functions.
These two topics appear to have nothing to do with each other whatsoever, and hence
the work of von Neumann (1932) in which they are related seems nothing short of a
miracle. The aim of my paper is to put this miracle in some historical perspective.
6
See Mehrtens (1990) and Gray (2008) in general, and Siegmund-Schultze (1982) for functional analysis.
7
‘One cannot overstate the significance of the influence exerted by Hilbert’s thought and personality on
all who came out of [the Mathematical Institute at Göttingen] ’ (Corry, 2018). See also Rowe (2018).
8
See e.g. Corry (1997, 2004a, 2018) and references therein. It is puzzling that Hilbert did not mention
Newton’s Principia in this light, which was surely the first explicit and successful axiomatization of physics.
3
2 Hilbert: Axiomatic method
I believe this: as soon as it is ripe for theory building, anything that can be the subject
of scientific thought at all falls under the scope of the axiomatic method and hence
indirectly of mathematics. By penetrating into ever deeper layers of axioms in the
sense outlined erlier we also gain insight into the nature of scientific thought by itself
and become steadily more aware of the unity of our knowledge. Under the header of
the axiomatic method mathematics appears to be called into a leading role in science
in general. (Hilbert, 1918, p. 115).
Hilbert (1918) begins his essay on axiomatic thought (of which the above text is the end)
by stressing the importance of the connection between mathematics and neigbouring fields,
especially physics and epistemology, and then says that the essence of this connection lies
in the axiomatic method. By this, he simply means the identification of certain sentences
(playing the role of axioms) that form the foundation of a specific field in the sense that
its theoretical structure (Hilbert uses the German word Fachwerk ) can be (re)constructed
from the axioms via logical principles. Axioms typically state relations between “things”
(Dinge), like “points” or “lines”, which are defined implicitly through the axioms and
hence may change their meaning if the axiom systems in which they occur change, as is
the case in e.g. non-Euclidean geometry.9 The epistemological status of the axioms differs
between fields. For example, Hilbert considered geometry initially a natural science:
Geometry also emerges from the observation of nature, from experience. To this ex-
tent, it is an experimental science. (. . . ) All that is needed is to derive [its] foundations
from a minimal set of independent axioms and thus to construct the whole edifice of
geometry by purely logical means. In this way geometry is turned into a purely math-
ematical science. (Hilbert in 1898–99, quoted in Corry, 2004a, p. 90).10
This does not mean that he treated the axioms of geometry as “true” (as Euclid had
done): Hilbert often stressed the tentative and malleable nature of axiom systems,11 and
acknowledged that axioms for physics might even be inconsistent, in which case finding
new, consistent axioms is an important source of progress (Corry, 2004a; Majer, 2014).12
Hilbert is famous for his purely formal treatment of axioms,13 which indeed was striking
all the way from the Grundlagen der Geometrie in 1899 to his swan song Grundlagen
der Mathematik (Hilbert & Bernays, 1934, 1939), but in fact such formality is always
strictly limited to the logical analysis of axiom systems (notably his relentless emphasis
on consistency and to a lesser extent on completeness) and the validation of proofs. Indeed,
except for logic Hilbert made almost no contribution to the axiomatization of mathematical
structures, although, starting already in the 19th century with e.g. Dedekind, Peano and
Weber, this became a central driving force of 20th century mathematics (Corry, 2004b).
9
The revolutionary nature of this view may be traced from Hilbert’s correspondence with Frege, who
apparently never accepted (or even grasped) this point (Gabriel et al, 1980; Blanchette, 2018).
10
From unpublished lecture notes by Hilbert, emphasis in original. Translation: Corry.
11
As exemplified by the seven editions of Grundlagen der Geometrie Hilbert published during his lifetime!
12
Though Einstein would speak of “principles” rather than “axioms”, many of his key contributions to
physics, such as special relativity, general relativity, and the EPR-argument are examples of this strategy.
13
This has earned Hilbert the undeserved reputation of being a “formalist”, which is remote from his
actual views on mathematics. The purely symbolic treatment of axioms and proofs was not even new with
Hilbert; he apparently took it from Russell (Mancosu, 2003; Ewald & Sieg, 2013) and hence indirectly
from Frege and Peano. But unlike Russell, until the last decade of his career dedicated to proof theory,
Hilbert stated axioms quite informally, using a combination of mathematical and natural language.
4
3 Hilbert: Functional analysis
Inspired by Fredholm’s theory of integral operators,14 in the paper mentioned by Dieudonné
in the Introduction above, Hilbert (1906) introduced many of the key tools of functional
analysis, such as bounded and compact operators and spectral theory, culminating in his
discovery of continuous spectra.15 However, what we now see as the central aspect of
functional analysis, namely its linear structure,
P is absent! Hilbert’s analysis is entirely
given in terms of quadratic forms K(x) = p,q kpq xp xq , where the sequence (x) satsifies
P 2
k |x|k ≤ 1, so that he works on what we would now call the closed unit ball of the Hilbert
space ℓ2 of square-summable sequences (which is compact in what we now call the weak
topology, which Hilbert also introduced himself and heavily exploited).
As pointed out at the end of §2, though at first sight odd,16 it seems typical for Hilbert
not to rely on abstract axiomatized mathematical structures, let alone that he would care
to refer to Peano (1888), in which the concepts of a vector space and a linear map had
been axiomatized. Perhaps Peano’s axiomatization was really unknown in Göttingen,
where Hilbert’s former student Weyl (1918) rediscovered the axioms for a vector space
in the context of Einstein’s theory of General Relativity (at least, he did not cite Peano
either).17 However, the essentially linear nature of Hilbert’s constructions was soon noted
and developed by various mathematicians, notably Hilbert’s own student Schmidt (1908),
who introduced ℓ2 including its inner product and even norm in modern form,18 and a
bit later by Riesz (1913), who rewrote most of Hilbert’s results in the modern way via
bounded or compact linear operators on Schmidt’s space ℓ2 (the notion of a linear operator
as such had already appeared before, notably in the Italian school of functional analysis).
Around 1905, Hadamard and his student Fréchet (partly inspired by the Italian school)
emphasised the idea of looking at functions as points in some (infinite-dimensional) vector
space, including an early use of topology, then also a new field–it was Fréchet (1906) who
in his thesis introduced metric spaces. This idea, often seen as the essence of functional
analysis, crossed the Hilbert school through the introduction of L2 -spaces, including the
spectacular and unexpected isomorphism L2 ([a, b]) ∼ = ℓ2 due to Riesz (1907) and Fischer
(1907). A truly geometric or spatial view of functional analysis was subsequently developed
especially by Riesz (1913, 1918), culminating in the axiomatic development of Banach
spaces in the 1920s by Helly, Wiener, and Banach (Monna, 1973; Pietsch, 2007).19
14
[The day (in 1901) on which Holmgren spoke on Fredholm’s work in Hilbert’s seminar] ‘was decisive
for a long period in Hilbert’s life and for a considerable part of his fame’ (Blumenthal, 1935, p. 410).
15
Since he lacked the concept of a linear operator Hilbert used a somewhat cumbersome definition of a
spectrum; the modern definition is due to Riesz (1913) and was also adopted by von Neumann, see §4.
16
Dieudonné (1981) explains the 19th century emphasis on matrices and quadratic forms at the expense
of vectors and linear maps, so that Hilbert had one foot in the 19th century and the other in the 20th.
17
See, however, Corry (2004a, §9.2) on the culture of “nostrification” in Hilbert’s Göttingen: ‘It was
widely understood, among German mathematicians at least, that “nostrification” encapsulated the peculiar
style of creating and developing scientific ideas in Göttingen, and not least because of the pervasive influence
of Hilbert. Of course, “nostrification” should not be understood as mere plagiarism.’ (loc. cit. p. 419).
18
The first author to use the term “Hilbert space” (Hilbertscher Raum) was Schönflies (1908), but he
meant the closed unit ball in ℓ2 (which was historically spot on, since that is what Hilbert analysed!).
Riesz (1913) used l’espace hilbertien for what we now call ℓ2 , and both the notion and the name Hilbert
space for the general abstract concept we now take it to mean was introduced by von Neumann (1927a).
19
As an intermediate step from Hilbert to Banach spaces, Lp spaces were introduced by Riesz (1909).
Many historians point out that (Frigyes) Riesz was familiar with the Italian, French, and German schools
of functional analysis. Dieudonné (1981, p. 145) calls Riesz (1918), which develops the (spectral) theory
of compact operators on Banach spaces avant la lettre, ‘one of the most beautiful papers ever written.’
5
4 von Neumann: Foundations of quantum theory
Methoden der mathematischen Physik by Courant and Hilbert (1924) put the lid on the
Göttingen school in functional analysis. It was meant to mathematize classical physics.
And now, one of those events happened, unforeseeable by the wildest imagination,
the like of which could tempt one to believe in a pre-established harmony between
physical nature and mathematical mind:20 Twenty years after Hilbert’s investigations
quantum mechanics found that the observables of a physical system are represented
by the linear symmetric operators on a Hilbert space and that the eigen-values and
eigen-vectors of that operator which represent energy are the energy levels and corre-
sponding stationary quantum states of the system. Of course, this quantum-physical
interpretation added greatly to the interest in the theory and led to a more scrupulous
investigation of it, resulting in various simplifications and extension. (Weyl, 1951, p.
541).
Although Weyl (who more often fell into lyrical overstatements in his philosophical writ-
ings) may have been right about the events in question being unforeseeable, the historical
record shows considerable continuity, too. Perhaps a slightly more balanced judgement is:
This revolution was made possible by combining a concern for rigorous foundations
with an interest in physical applications, and by coordinating the relevant literature
in depth. (Birkhoff & Kreyszig, 1984, pp. 306–307).
At least the first two aspects were exemplified by Hilbert, who had lectured on the math-
ematical foundations of physics since 1898 (Sauer& Majer, 2009; Majer & Sauer, 2021),
and, helped by various assistants,21 organised a regular research seminar on the latest
developments in physics (Reid, 1970; Schirrmacher, 2019). With Born, Heisenberg, and
Jordan all at Göttingen at the time, during the Winter Semester of 1926–1927 Hilbert lec-
tured on quantum theory, with book-length lecture notes by Nordheim.22 These lectures
are very impressive and cover almost everything, from Hamilton–Jacobi theory and the
“old quantum theory” to Heisenberg’s matrix mechanics, Schrödinger’s wave mechanics,
Born’s probability interpretation, and finally Jordan’s Neue Begründung, i.e. his attempt
(simultaneous with Dirac’s) to unify the last three ingredients into a single formalism.23
‘Coordinating the relevant literature in depth’ was therefore certainly taken care of on
the physics side, but on the mathematical side Hilbert’s surprising lack of interest in the
axiomatisation of new mathematical structures except logic (cf. §2) still played a role:
The German school [in functional analysis, i.e. Hilbert’s school] remained reserved
with respect to the more abstract concepts of set theory and axiomatics until well into
the 1920s. (Siegmund-Schultze, 2003, p. 385).
20
“Pre-established harmony”, a philosophical concept originally going back to Leibniz, was a popular
concept in the Göttingen of Hilbert, where it referred to the relationship between mathematics and physics,
or more generally between the human mind and nature (Pyenson, 1982; Corry, 2004a). Minkowski, Hilbert,
Born, and Weyl himself all used it as approriate, and Corry (2004a, pp. 393–394) even claims that it was
‘one of the most basic concepts that underlay the whole scientific enterprise in Göttingen’, adding that
‘Hilbert, like all his colleagues in Göttingen, was never really able to explain, in coherent philosophical
terms, its meaning and the possible basis of its putative pervasiveness, except by alluding to “a miracle”.’
21
Hilbert’s first assistant (at the time unpaid) had been Born in 1904; from 1922–1926 it was Nordheim.
22
These may be found in Sauer& Majer (2009), pp. 507–706. Half of the course, on the “old quantum
theory” was practically reproduced from Hilbert’s earlier lectures on quantum theory during 1922–1923.
23
See Duncan & Janssen (2009, 2013) for a detailed survey of Jordan’s Neue Begründung. Older and
somewhat complementary histories of this period include Jammer (1989) and Mehra & Rechenberg (2000).
6
By a simple twist of fate, in 1926 Hilbert attracted the internationally acknowledged young
genius von Neumann to spend the academic year 1926–1927 to Göttingen in order to work
on his Proof Theory,24 but in actual fact the latter mostly worked on the mathematical
foundations of quantum theory and thus filled in the abstraction and axiomatisation gap.25
The paper by Hilbert, von Neumann, & Nordheim (1927), now obsolete, is actually a
summary of Hilbert’s (inconclusive) views, based on his lectures; the decisive establish-
ment of the interaction between quantum theory and functional analysis is entirely due
to von Neumann (1927ab), with mathematical details further elaborated in von Neumann
(1930ab) and a unified exposition in his famous book from 1932, which remains a classic.
Apart from his discussion of quantum statistical mechanics and of the measurement prob-
lem, which were path-breaking contributions to physics but are less relevant for our topic,
the main accomplishments of von Neumann (1932) in the light of functional analysis are:26
1. Axiomatisation of the notion of a Hilbert space (previously known only in examples).
2. Establishment of a spectral theorem for (possibly unbounded) self-adjoint operators.
3. Axiomatisation of quantum mechanics in terms of Hilbert spaces (and operators):
(a) Identification of observables with (possibly unbounded) self-adjoint operators.
(b) Identification of pure states with one-dimensional projections (or rays).
(c) Identification of transition amplitudes with inner products.
(d) A formula for the Born rule stating the probability of measurement outcomes.
(e) Identification of general states with density operators.
(f) Identification of propositions with closed subspaces (or the projections thereon).
In particular, von Neumann provided two separate (but closely related) axiomatisations:27
• of Hilbert space as an abstract mathematical structure (almost contra Hilbert);
• of quantum mechanics as a physical theory (entirely in the spirit of Hilbert).
Ad 1. Although Schmidt, Riesz, and others, had thought about sequence and function
spaces like ℓ2 and L2 in a geometric way 20 years earlier, including the use of inner
products, orthogonality, and norms, the abstract concept of a Hilbert space (unlike that
of a Banach space) was still lacking before 1927. The novelty of von Neumann’s coordinate-
free approach to Hilbert spaces is illustrated by the fatherly advice Schmidt gave him:
No! No! You shouldn’t say operator, say matrix! (Bernkopf, 1967, p. 346).
Ad 2. This was a vast abstraction and generalisation of practically all of the spectral theory
done in Hilbert’s school, including the work of Weyl and Carleman on what (since von
Neumann) are called unbounded self-adjoint operators and their deficiency indices.28
24
Hilbert got a fellowship for von Neumann from the International Education Board (a subsidiary of
the Rockefeller Foundation). In the Fall of 1927 von Neumann moved to Berlin as a Privatdocent, where
Hilbert’s former student Schmidt provided him with the concept of self-adjointness he had initially missed
in setting up his spectral theory for closed unbounded operators (Birkhoff & Kreyszig, 1984, p. 309).
25
von Neumann had made brilliant contributions to set theory already in his late teens. For further
information about von Neumann see Oxtoby et al (1958), Heims (1980), Glimm et al (1990), Macrae (1992),
Bródy & Vámos (1995), Rédei (2005b), and the rare but insightful manuscript Vonneumann (1987).
26
Even finding one of these would have been impressive, not only for someone who was 23 years old.
27
See Lacki (2000), Rédei (2005a), and Rédei & Stöltzner (2006) on von Neumann’s methodology.
28
This history is very nicely explained by Dieudonné (1981), Chapter VII. See also Stone (1932).
7
Ad 3. This remains the basis for any discussion of the foundations of quantum theory.
Ad 3(a). One of von Neumann’s main goals was a rigorous proof of the equivalence be-
tween matrix mechanics (which almost deliberately lacked states) and wave mechanics
(which initially lacked observables). As a first ingredient, Heisenberg’s matrices (which
by themselves were ‘quantum-mechanical reinterpretations of classical observables’) were
reinterpreted once again, now as self-adjoint operators on some Hilbert space like ℓ2 . The
need for unbounded operators, e.g. for position, momentum, and energy, emerged at once.
Ad 3(b). And this was the second ingredient of the equivalence proof. Identifying Schröding-
er’s wave-function Ψ with a unit vector in the Hilbert space L2 (R3 ) was an accomplishment
by itself, but on top of this, von Neumann (and Weyl) quickly recognized the importance
of the fact that such vectors only define (pure) states up to a phase. Thus the cleanest
way to define (pure) states is to identify them with 1d projections rather than vectors.29
Ad 3(c). Next to the goal just stated, the point of von Neumann’s axiomatisation was to
provide a home to the mysterious transition amplitudes hϕ|ψi that were at the heart of
Jordan’s Neue Begründung (Duncan & Janssen, 2013), which also Born, Dirac, and Pauli
regarded as the essence of quantum mechanics. If |ψi and |ϕi are unit vectors in some
Hilbert space, von Neumann took the amplitude hϕ|ψi to be their inner product, with
corresponding transition probability P (ϕ, ψ) = |hϕ|ψi|2 . In terms of the one-dimensional
projections e = |ϕihϕ| and f = |ψihψ|, this gives P (e, f ) = Tr (ef ), where Tr is the trace.30
Ad 3(d). Von Neumann’s Born probability to find a result λ ∈ I in a measurement of some
self-adjoint operators A in a state ρ is given by Tr (ρE(I)), where E(I) is the spectral
projection for a subset I in the spectrum of A (and generalisations thereof to commuting
observables). For ρ = |ψihψ|, I = {λ}, and E(I) = |ϕihϕ|, assuming Aϕ = λϕ and the
eigenvalue λ is nondegenerate, this recovers the transition probability in the previous item.
Ad 3(e). Von Neumann tried to prove this identification by showing that if Exp is a linear
map from the (real) vector space of all bounded self-adjoint operators A on some Hilbert
space H to R that is normalized (Exp(I)) = 1), dispersion-free (Exp(A2 ) = Exp(A)2 ),
and satisfies a continuity condition (which is automatic if H is finite-dimensional), then
Exp(A) = Tr (ρA) for some density operator ρ on H. Unfortunately, he mistook this
correct, non-circular, and interesting result for a proof that no hidden variables can underly
quantum mechanics. See Bub (2011) and Dieks (2016) for a fair and balanced account.
Ad 3(f ). This was further developed by Birkhoff & von Neumann (1936), whose lattice-
theoretic calculus of such propositions initiated the field of quantum logic (Rédei, 1998).
29
Following Minkowski’s example from his Geometry of Numbers, in the completely different context of
quantum mechanics von Neumann defined pure states to be extreme points of the convex set of all states.
30
It should be admitted that this did not clinch the issue. Dirac and Jordan used probability amplitudes
like hx|pi = exp(−ixp/~), where x, p ∈ R, but “eigenstates” like |xi and |pi for the continuous spectrum
of some operator (like position and momentum here) are undefined in Hilbert space and hence have no
inner product. Von Neumann circumvented this problem by first practically starting his book with a tirade
against Dirac’s mathematics, and second, by writing down expressions like Tr (E(I)F (J)), where E(I) and
F (J) are the spectral projections for subsets I and J in the spectra of some self-adjoint operators A and
B, respectively. Unfortunately, these “transition probabilities” are not only unnormalized; they may even
be infinite. This was one of the reasons why von Neumann probably felt uncomfortable with his formalism
right from the start, and later sought a way out of this problem through a combination of lattice theory à
la Birkhoff & von Neumann (1936) and the theory of operator algebras he had also created himself, cf. §5.2.
The key is the existence of type ii1 factors, which admit a normalised trace tr, i.e. tr(I) = 1. Replacing
Tr (E(I)F (J)) by tr(E(I)F (J)) then makes the transition probabilities finite as well as normalized, but
this only works if the spectral projections lie in the said factor. See Araki (1990) and Rédei (1996, 2001).
8
5 Quantum theory and functional analysis since 1932
In this section we give a brief overview of three areas of functional analysis that have had
fruitful interactions with quantum theory since the initial breakthrough during 1927–1932.
9
given Hilbert space is replaced by an arbitrary (abstract) operator algebra. As such,
one may continue to work with states, observables, and expectation values (Segal, 1947).
From the 1960s onwards operator algebras have become an important tool in mathematical
physics, initially applied to quantum systems with infinitely many degrees of freedom, as in
quantum statistical mechanics (Ruelle, 1969; Bratteli & Robinson, 1981, 1987; Haag, 1992;
Simon, 1993) and quantum field theory (Haag, 1992; Araki, 1999; Brunetti, Dappiaggi, &
Fredenhagen, 2015).36 In turn, these physical applications have also given rise to various
new mathematical ideas.37 Furthermore, since the 1980s the field of operator algebras has
been greatly refined and expanded by the toolkit of noncommutative geometry (Connes,
1994), which so far has been applied to many areas of physics, ranging from particle
physics (Connes & Marcolli, 2008; van Suijlekom, 2015) to deformation quantization and
the classical limit of quantum mechanics (Rieffel, 1994; Landsman, 1998).
5.3 Distributions
Another major development in functional analysis that is relevant for quantum physics
was the theory of distributions due to Schwartz (1950–1951).38 Though closely related
to Hilbert and Banach spaces through all kinds of natural embeddings and dualities,
spaces of distributions belong to the wider class of locally convex topological vector spaces,
which incidentally were introduced by von Neumann (1935). The facts that the Dirac
delta-function, which had annoyed von Neumann (1932, p. 2) so much, becomes a well-
defined object in distribution theory, and that the “rigged Hilbert space” approach to
distributions (Gelfand & Vilenkin, 1964) even gives a rigorous and satisfactory version of
Dirac’s continuous eigenfunction expansions (Maurin, 1968), have had surprisingly little
impact on quantum mechanics, even when it is seen in the light of mathematical physics.39
Instead, the main applications of distribution theory to quantum physics have been to
quantum field theory, in at least three originally different but closely related ways:
1. Through Wightman’s Axiomatic Quantum Field Theory, where quantum fields are
defined as unbounded operator-valued distributions (Streater & Wightman, 1964).
10
Epilogue
Functional analysis arose in the early twentieth century and gradually, conquering one
stronghold after another, became a nearly universal mathematical doctrine, not merely
a new area of mathematics, but a new mathematical world view. Its appearance was
the inevitable consequence of the evolution of all of nineteenth-century mathematics,
in particular classical analysis and mathematical physics. (. . . ) Its existence answered
the question of how to state general principles of a broadly interpreted analysis in a way
suitable for the most diverse situations. (Vershik, 2006, p. 438, quoted by MacCluer,
2009, p. vii).
This passage explains to some extent why the spectacular and unexpected application
of functional analysis to quantum theory was possible: though originating in problems
from classical physics, the “modernist” turn of mathematics towards abstraction and ax-
iomatisation that brought the subject into the 20th century made almost every field of
mathematics universally applicable. Moreover, much as quantum theory was originally
meant to merely describe the atomic domain but subsequently, through its extension to
quantum field theory in fact turned out to be a theory of all of physics (except perhaps
gravity), through von Neumann’s invention of operator algebras as well as Schwartz’s
theory of distributions (both partly inspired by quantum mechanics), functional analysis
continued to provide an appropriate mathematical language also for quantum field theory.
Having said this, the question why functional analysis (here taken to be the original lin-
ear theory) and especially Hilbert spaces (or operator algebras) underlie quantum physics
remains unanswered. Perhaps starting with Birkhoff & von Neumann (1936), many people
have tried to derive the mathematical formalism from plausible physical principles, but
I believe that every such derivation so far contains a contingent or even incomprehensi-
ble part in order to derive the (complex) Hilbert space formalism.42 In this respect, the
connection between quantum theory and functional analysis remains mysterious.
Finally, let me note that this was a winner’s (or “whig”) history, full of hero-worship:
following in the footsteps of Hilbert, von Neumann established the link between quantum
theory and functional analysis that has lasted. Moreover, partly through von Neumann’s
own contributions (which are on a par with those of Bohr, Einstein, and Schrödinger), the
precision that functional analysis has brought to quantum theory has greatly benefited the
foundational debate. However, it is simultaneously a loser’s history: starting with Dirac
and continuing with Feynman, until the present day physicists have managed to bring
quantum theory forward in utter (and, in my view, arrogant) disregard for the relevant
mathematical literature. As such, functional analysis has so far failed to make any real
contribution to quantum theory as a branch of physics (as opposed to mathematics), and
in this respect its role seems to have been limited to something like classical music or other
parts of human culture that adorn life but do not change the economy or save the planet.
On the other hand, like General Relativity, perhaps the intellectual development reviewed
in this paper is one of those human achievements that make the planet worth saving.
42
In Birkhoff & von Neumann (1936) the modular law is already problematic; in refinements of their
lattice-theoretic approach based on the reconstruction theorem of Solèr (1995) one has to assume ortho-
modularity and the existence of an infinite orthonormal set (and still needs further arguments to single
out C over R or H), etc. Mackey (1963) himself admits defeat by simply postulating that the lattice of
propositions of a quantum system, for which he first gives many promising axioms, is isomorphic to the
projection lattice P(H) of some Hilbert space H. In my own approach based on the axiomatisation of
transition probability spaces, axiom C ∗ 2 on page 104 of Landsman (1998), which prescribes the transition
probabilities of a 2-level system, seems to lack any physical justification. See also Grinbaum (2007).
11
References
[1] Araki, H. (1990). Some of the legacy of John von Neumann in physics: Theory of measurement,
quantum logic and von Neumann algebras in physics. In: Glimm et al (1990), pp. 119–136.
[2] Araki, H. (1999). Mathematical Theory of Quantum Fields. Oxford: Oxford University Press.
[3] Banach, S. (1932). Théorie des Opérations Linéaires. Warszawa: Instytut Matematyczny
Polskiej Akademii Nauk, New York: Chelsea.
[4] Bär, C., Fredenhagen, K., eds. (2009). Quantum Field Theory on Curved Spacetimes: Concepts
and Mathematical Foundations. Berlin: Springer.
[5] Bernkopf, M.(1966). The development of function spaces with particular reference to their
origins in integral equation theory. Archive for History of Exact Sciences 3, 1–96.
[6] Birkhoff, G. & Kreyszig, E. (1984). The establishment of functional analysis. Historia Math-
ematica 11, 258–321.
[7] Birkhoff, G., von Neumann, J. (1936). The logic of quantum mechanics. Annals of Mathematics
37, 823–843.
[8] Blanchette, P. (2018). The Frege–Hilbert Controversy. The Stanford Encyclopedia of Philoso-
phy (Fall 2018 Edition), Zalta, E.N. (ed.).
https://plato.stanford.edu/archives/fall2018/entries/frege-hilbert/.
[9] Blumenthal, O. (1935). Lebensgeschichte. David Hilbert: Gesammelte Abhandlungen Vol. III,
pp. 388–429. Berlin: Springer.
[10] Bohm, A. (1994). Quantum Mechanics: Foundations and Applications. New York: Springer.
[11] Bratteli, O., Robinson, D.W. (1987). Operator Algebras and Quantum Statistical Mechanics.
Vol. I: C*- and W*-Algebras, Symmetry Groups, Decomposition of States. Second Edition.
Berlin: Springer.
[12] Bratteli, O., Robinson, D.W. (1981). Operator Algebras and Quantum Statistical Mechanics.
Vol. II: Equilibrium States, Models in Statistical Mechanics. Berlin: Springer.
[13] Bródy, F., Vámos, eds. (1995). The Neumann Compendium. Singapore: World Scientific.
[14] Brunetti, R., Dappiaggi, C., Fredenhagen, K., eds. (2015). Advances in Algebraic Quantum
Field Theory. Dordrecht: Springer.
[15] Bub, J. (2011). Is von Neumann’s ‘no hidden variables’ proof silly?, Deep Beauty: Mathe-
matical Innovation and the Search for Underlying Intelligibility in the Quantum World, pp.
393–408. Halvorson, H., ed. Cambridge: Cambridge University Press.
[16] Connes, A. (1994). Noncommutative Geometry. San Diego: Academic Press.
[17] Connes, A., Marcolli, M. (2008). Noncommutative Geometry, Quantum Fields, and Motives.
New Delhi: Hindustan Book Agency.
[18] Corry, L. (1997). David Hilbert and the axiomatization of physics (1894–1905). Archive for
History of Exact Sciences 51, 83–198.
[19] Corry, L. (2004a). David Hilbert and the Axiomatization of Physics (1898–1918): From Grund-
lagen der Geometrie to Grundlagen der Physik. Dordrecht: Kluwer.
[20] Corry, L. (2004b). Modern Algebra and the Rise of Mathematical Structures. Second revised
edition. Basel: Springer.
[21] Corry, L. (2018). Hilbert’s sixth problem: between the foundations of geometry and
the axiomatization of physics. Philosophical Transactions of the Royal Society A. DOI:
10.1098/rsta.2017.0221.
12
[22] Courant, R., Hilbert, D. (1924). Methoden der mathematischen Physik I. Berlin: Springer.
[23] Dieks, D. (2016). Von Neumann’s impossibility proof: Mathematics in the service of rhetorics.
Studies in History and Philosophy of Modern Physics 60, 136–148.
[24] Dieudonné, J. (1981). History of Functional Analysis. Amsterdam: North-Holland.
[25] Dijksterhuis, E.J. (1961). The Mechanisation of the World Picture. Oxford: Oxford University
Press. Translation of De Mechanisering van het Wereldbeeld (Meulenhoff, Amsterdam, 1950).
[26] Dirac, P.A.M. (1930). The Principles of Quantum Mechanics. Oxford: Clarendon Press.
[27] Doran, R.S., ed. (1994). C*-algebras: 1943–1993. Contemporary Mathematics Vol. 167. Prov-
idence: American Mathematical Society.
[28] Doran, R.S., Belfi, V. (1986). Characterization of C*-algebras. New York: Marcel Dekker.
[29] Dorier, J.-L. (1995). A general outline of the genesis of vector space theory. Historia Mathe-
matica 22, 227–261.
[30] Duncan, A., Janssen, M. (2009). From canonical transformations to transformation theory,
1926–1927: The road to Jordans Neue Begründung. Studies in History and Philosophy of
Modern Physics 40, 352–362.
[31] Duncan, A., Janssen, M. (2013). (Never) Mind your p’s and q’s: Von Neumann versus Jordan
on the foundations of quantum theory. The European Physical Journal H 38, 175–259.
[32] Eijndhoven, S.J.L. van, Graaf, J. de (1986). A Mathematical Introduction to Dirac’s Formal-
ism. Amsterdam: North-Holland (Elsevier).
[33] Epstein, H., Glaser, V. (1973). The role of locality in perturbation theory. Annales de
l’Institute Henri Poincaré A (Physique théorique) 19, 211–295.
[34] Ewald, W., Sieg, W. (2013). David Hilbert’s Lectures on the Foundations of Arithmetic and
Logic 1917-1933. Heidelberg: Springer.
[35] Fischer, E. (1907). Sur la convergence en moyenne. Comptes Rendus de l’Académie des Sci-
ences 144, 1022–1024.
[36] Fréchet, N. (1906). Sur quelques points du calcul functionel. Rendiconti del Circolo Matematico
di Palermo 22, 1–74.
[37] Gabriel, G., Kambartel, F., Thiel, C. (1980). Gottlob Frege’s Briefwechsel mit D. Hilbert, E.
Husserl, B. Russell, sowie ausgewählte Einzelbriefe Freges. Hamburg: Felix Meiner Verlag.
[38] Gelfand, I.M., Naimark, M.A. (1943). On the imbedding of normed rings into the ring of
operators in Hilbert space. Sbornik: Mathematics 12, 197–213.
[39] Gelfand, I.M., Vilenkin, N.J. (1964). Generalized Functions. Vol. 4: Some Applications of
Harmonic Analysis. Rigged Hilbert Spaces. New York: Academic Press.
[40] Gérard, C. (2019). Microlocal analysis of quantum fields on curved spacetimes.
https://arxiv.org/abs/1901.10175.
[41] Glimm, J., Impagliazzo, J., Singer, I., eds. (1990). The Legacy of John von Neumann. Pro-
ceedings of Symposia in Pure Mathematics 50. Providence: American Mathematical Society.
[42] Gray, J.D. (2008). Plato’s Ghost: The Modernist Transformation of Mathematics. Princeton:
Princeton University Press.
[43] Grinbaum, A. (2007). Reconstruction of quantum theory. British Journal for the Philosophy
of Science 58, 387–408.
[44] Haag, R. (1992). Local Quantum Physics: Fields, Particles, Algebras. Heidelberg: Springer.
[45] Haag, R. (2010). Some people and some problems met in half a century of commitment to
mathematical physics. The European Physics Journal H 35, 263–307.
13
[46] Heims, S.J. (1980). John von Neumann and Norbert Wiener: From Mathematics to the Tech-
nologies of Life and Death. Cambridge (Mass.): MIT Press.
[47] Hilbert, D. (1902). Mathematical Problems. Lecture delivered before the International
Congress of Mathematicians at Paris in 1900. Bulletin of the American Mathematical Society
8, 437–479. Translated from Göttinger Nachrichten, 1900, pp. 253–297.
[48] Hilbert, D. (1906). Grundzüge einer allgemeinen Theorie der linearen Integralgleichun-
gen. Vierte Mitteilung. Nachrichten von der Gesellschaft der Wissenschaften zu Göttingen,
Mathematisch-Physikalische Klasse, 157–227. Reprinted in Hilbert (1912).
[49] Hilbert, D. (1912). Grundzüge einer allgemeinen Theorie der linearen Integralgleichungen.
Leipiz und Berlin: Teubner. https://archive.org/details/grundzugeallg00hilbrich.
[50] Hilbert, D. (1918). Axiomatisches Denken. Mathematische Annalen 78, 405–415. Reprinted
in David Hilbert: Gesammelte Abhandlungen Vol. III, pp. 146–156 (1935). Berlin: Springer.
[51] Hilbert, D., Bernays, P. (1934, 1939). Grundlagen der Mathematik, Bd. I, II. Berlin: Springer.
[52] Hilbert, D., von Neumann, J., Nordheim, L. (1927). Über die Grundlagen der Quanten-
mechanik. Mathematische Annalen 98, 1–30.
[53] Hörmander, L. (1990). The Analysis of Linear Partial Differential Operators I. Second Edition.
Berlin: Springer.
[54] Jammer, M. (1989). The Conceptual Development of Quantum Mechanics. Second Edition.
New York: American Institute of Physics.
[55] Jordan, P., von Neumann, J., Wigner, E.P. (1934). On an algebraic generalization of the
quantum mechanical formalism. Annals of Mathematics 35, 29–64.
[56] Kadison, R.V. (1982). Operator algebras: The first forty years. Proceedings of Symposia in
Pure Mathematics 38(1), pp. 1–18. Providence: American Mathematical Society.
[57] Kato, T. (1951). Fundamental properties of Hamiltonian operators of Schrödinger type. Trans-
actions of the American Mathematics Society 70, 195–211.
[58] Kato, T. (1966). Perturbation Theory for Linear Operators. Berlin: Springer.
[59] Lacki, J. (2000). The early axiomatizations of quantum mechanics: Jordan, von Neumann
and the continuation of Hilbert’s program. Archive for History of Exact Sciences 54, 279–318.
[60] Landsman, K. (1998). Mathematical Topics Between Classical and Quantum Mechanics. New
York: Springer.
[61] Landsman, K. (2017). Foundations of Quantum Theory: From Classical Concepts to Operator
Algebras. Cham: Springer Open. https://www.springer.com/gp/book/9783319517766.
[62] MacCluer, B.D. (2009). Elementary Functional Analysis. New York: Springer.
[63] Mackey, G.W. (1963). The Mathematical Foundations of Quantum Mechanics. New York:
Benjamin.
[64] Macrae, N. (1992). John von Neumann: The Scientific Genius Who Pioneered the Modern
Computer, Game Theory, Nuclear Deterrence, and Much More. Providence: American Math-
ematical Society.
[65] Majer, U. (2014). The “axiomatic method” and its constitutive role in physics. Perspectives
on Science 22, 56–79.
[66] Majer, U., Sauer, T., Eds. (2021). David Hilbert’s Lectures on the Foundations of Physics
1898–1914.
[67] Mancosu, P. (2003). The Russellian influence on Hilbert and his school. Synthese 137, 59–101.
14
[68] Maurin, K. (1968). Generalized Eigenfunction Expansions and Unitary Representations of
Topological Groups. Warsaw: Polish Scientific Publishers.
[69] Mehra, J., Rechenberg, H. (2000). The Historical Development of Quantum Theory. Vol. 6:
The Completion of Quantum Mechanics 1926–1941. Part 1: The Probabilistic Interpretation
and the Empirical and Mathematical Foundation of Quantum Mechanics, 1926-1936. New
York: Springer-Verlag.
[70] Mehrtens, H. (1990). Moderne Sprache Mathematik. Frankfurt a/M: Suhrkamp.
[71] Monna, A.F. (1973). Functional Analysis in Historical Perspective. Utrecht: Oosthoek.
[72] Monna, A.F. (1973). Dirichlet’s Principle: A mathematical Comedy of Errors and its Influence
on the Development of Analysis. Utrecht: Oosthoek, Scheltema and Holkema.
[73] Moore, G.H. (1995). The axiomatization of linear algebra: 1875–1940. Historia Mathematica
22, 262–303.
[74] Neumann, J. von (1927a). Mathematische Begründung der Quantenmechanik. Nachrichten
von der Gesellschaft der Wissenschaften zu Göttingen, Mathematisch-Physikalische Klasse,
1–57. https://eudml.org/doc/59215.
[75] Neumann, J. von (1927b). Wahrscheinlichkeitstheoretischer Aufbau der Quantenmechanik.
ibid. 245–272. https://eudml.org/doc/59230.
[76] Neumann, J. von (1930a). Allgemeine Eigenwerttheorie hermitischer Funktionaloperatoren.
Mathematische Annalen 102, 49–131.
[77] Neumann, J. von (1930b). Zur Algebra der Funktionaloperationen und Theorie der normalen
Operatoren. Mathematische Annalen 102, 370–427.
[78] Neumann, J. von (1931). Die Eindeutigkeit der Schröderingerschen Operatoren. Mathematis-
che Annalen 104, 570–578.
[79] Neumann, J. von (1932). Mathematische Grundlagen der Quantenmechanik. Berlin: Springer–
Verlag. English translation: Mathematical Foundations of Quantum Mechanics. Princeton:
Princeton University Press (1955).
[80] Neumann, J. von (1935). On complete topological spaces. Transactions of the American Math-
ematica Society 37, 1–20.
[81] Neumann, J. von (1936). On an algebraic generalization of the quantum mechanical formalism
(Part i). Sbornik: Mathematics (N.S.) 1, 415–484.
[82] Neumann, J. von (1961). Collected Works, Vol. III: Rings of Operators. Taub, A.H., ed.
Oxford: Pergamon Press.
[83] Oxtoby, J.C., Pettis, B.J., Price, G.B. (1958). John von Neumann: 1903–1957. Bulletin of the
American Mathematical Society 64, No. 3, Part 2.
[84] Peano, G. (1888). Calcolo geometrico secondo l’Ausdehnungslehre di H. Grassmann: preceduto
dalla operazioni della logica deduttiva. Torino: Fratelli Bocca Editori.
[85] Petz, D., Rédei, M. (1995). John von Neumann and the theory of operator algebras. The
Neumann Compendium, pp. 163–181. Bródy, F., Vámos, eds. Singapore: World Scientific.
[86] Pier, J. (2001). Mathematical Analysis During the 20th Century. New York: Oxford University
Press.
[87] Pietsch, W. (2007). History of Banach Spaces and Linear Operators. Basel: Birkhäuser.
[88] Pyenson, L.R. (1982). Relativity in late-Wilhelminian Germany: The appeal to a preestab-
lished harmony between mathematics and physics. Archive for History of Exact Sciences 27,
137-155.
15
[89] Rédei, M. (1996). Why John von Neumann did not like the Hilbert space formalism of quantum
mechanics (and what he liked instead). Studies in History and Philosophy of Modern Physics
27, 493–510.
[90] Rédei, M. (1998). Quantum Logic in Algebraic Approach. Dordrecht: Kluwer Academic Pub-
lishers.
[91] Rédei, M. (2001). Von Neumann’s concept of quantum logic and quantum probability. In:
Rédei & Stöltzner (2001), pp. 153–172.
[92] Rédei, M. (2005a). John von Neumann on mathematical and axiomatic physics. The Role of
Mathematics in the Physical Sciences, pp. 43–54. Boniolo, G. et al, eds. Dordrecht: Springer.
[93] Rédei, M. (2005b). John von Neumann: Selected Letters. Providence: American Mathememat-
ical Society.
[94] Rédei, M., Stöltzner, eds. (2001). John von Neumann and the Foundations of Quantum
Physics. Dordrecht: Kluwer.
[95] Rédei, M., Stöltzner (2006), Soft axiomatisation: John von Neumann on method and von
Neumann’s method in the physical sciences. Intuition and the Axiomatic Method, Carson, E.,
Huber, R. (eds.), pp. 235–250. Dordrecht: Springer.
[96] Reed, M., Simon, B. (1972–1978). Methods of Modern Mathematical Physics. Volume I: Func-
tional Analysis. Volume II: Fourier Analysis, Self-Adjointness. Volume III: Scattering Theory.
Volume IV: Analysis of Operators. New York: Academic Press.
[97] Reid, C. (1970). Hilbert. Berlin: Springer.
[98] Rejzner, K. (2016). Perturbative Algebraic Quantum Field Theory: An Introduction for Math-
ematicians. Cham: Springer.
[99] Rieffel, M.A. (1994). Quantization and C ∗ -algebras. Contemporary Mathematics 167, 66–97.
[100] Riesz, F. (1907). Sur les systèmes orthogonaux des functions. Comptes Rendus de l’Académie
des Sciences 144, 615–619.
[101] Riesz, F. (1909). Untersuchungen über Systeme integrierbarer Funktionen. Mathematische
Annalen 69, 449–497.
[102] Riesz, F. (1913). Les Systèmes d’Équations Linéaires à une Infinité d’Inconnues. Paris:
Gauthiers–Villars.
[103] Riesz, F. (1918). Über lineare Funktionalgleichungen. Acta Mathematica 41, 71–98.
[104] Rosenberg, J. (2004). A selective history of the Stone–von Neumann Theorem. Contemporary
Mathematics 365, 331353.
[105] Rowe, D.E. (2018). A Richer Picture of Mathematics: The Göttingen Tradition and Beyond.
Cham: Springer.
[106] Ruelle, D. (1969). Statistical Mechanics: Rigorous Results. New York: Benjamin.
[107] Sauer, T., Majer, U., Eds. (2009). David Hilbert’s Lectures on the Foundations of Physics
1915–1927. Dordrecht: Springer.
[108] Scharf, G. (1995). Finite Electrodynamics: The Causal Approach. Second Edition. Berlin:
Springer.
[109] Scharf, G. (2001). Quantum Gauge Theories: A True Ghost Story. New York: Wiley.
[110] Schirrmacher, A. (2019). Establishing Quantum Physics in Göttingen: David Hilbert, Max
Born, and Peter Debye in Context, 1900–1926. Cham: Springer.
[111] Schmidt, E. (1908). Über die Auflösung linearer Gleichungen mit unendlich vielen Unbekan-
nten. Rendiconti del Circolo Matematico di Palermo xxv, 53–77.
16
[112] Schönflies, A. (1908). Die Entwicklung der Lehre von den Punktmannigfaltigkeiten, Zweiter
Teil. Jahresberichte der deutschen Mathematiker-Vereinigung, Ergänzungsband. Leipzig: Teub-
ner.
[113] Schwartz, L. (1950–1951). Théorie des Distributions. Vols. 1 & 2. Paris: Hermann.
[114] Schwartz, L. (2001). A Mathematician Grappling with his Century. Basel: Birkhäuser.
[115] Segal, I.E. (1947). Postulates for general quantum mechanics. Annals of Mathematics 48,
930–948.
[116] Sieg, W. (2013). Hilbert’s Programs and Beyond. Oxford: Oxford University Press.
[117] Siegmund-Schultze, R. (1982). Die Anfänge der Funktionalanalysis und ihr Platz im Un-
wandlungsprozeß der Mathematik um 1900. Archive for History of Exact Sciences 26, 13–71.
[118] Siegmund-Schultze, R. (2003). The origins of functional analysis. A History of Analysis, pp.
385–408. Jahnke, H. N., ed. Providence: American Mathematical Society.
[119] Simon, B. (1993). The Statistical Mechanics of Lattice Gases. Vol. I. Princeton: Princeton
University Press.
[120] Simon, B. (2000). Schrödinger operators in the twentieth century. Journal of Mathematical
Physics 41, 3523–3555.
[121] Simon, B. (2018). Tosio Kato’s work on non-relativistic quantum mechanics, part 1 and part
2. Bulletin of Mathematical Sciences 8, 121-232 and ibid. 9, 1–99.
[122] Solèr, M.P. (1995). Characterization of Hilbert spaces by orthomodular spaces. Communica-
tions in Algebra 23, 219–243
[123] Steen, L.A. (1973). Highlights in the history of spectral theory. American Mathematical
Monthly 80, 359–381.
[124] Stone, M.H. (1931). Linear transformations in Hilbert space, III: Operational methods and
group theory. Proceedings of the National Academy of Sciences of the United States of America
16, 172–175.
[125] Stone, M.H. (1932). Linear Transformations in Hilbert space and their Applications to Anal-
ysis. Providence: American Mathematical Society.
[126] Stöltzner, M. (2001). Opportunistic axiomatics–von Neumann and the methodology of math-
ematical physics. In: Rédei & Stöltzner (2001), pp. 35–62.
[127] Streater, R.F., Wightman, A.S. (1964). PCT, Spin and Statistics, and All That. New York:
Benjamin.
[128] Suijlekom, W.D. van (2015). Noncommutative Geometry and Particle Physics. Dordrecht:
Springer.
[129] Summers, S.J. (2001). On the Stone–von Neumann uniqueness theorem and its ramifica-
tions. John von Neumann and the Foundations of Quantum Physics, pp. 135–152. Rédei, M.,
Stöltzner, eds. Dordrecht: Kluwer.
[130] Vershik, A.M. (2006). The life and fate of functional analysis in the twentieth century. Math-
ematical Events of the Twentieth Century, eds. Bolibruch, A.A. et al, pp. 437–447. Berlin:
Springer.
[131] Volkert, K. (2015). David Hilbert: Grundlagen der Geometrie. Berlin: Springer.
[132] Vonneumann, N. (1987). John von Neumann as Seen by his Brother. Typescript.
[133] Weyl, H. (1918). Raum - Zeit - Materie. Berlin: Springer.
[134] Weyl, H. (1928). Gruppentheorie und Quantenmechanik. Leipzig: Hirzel.
[135] Weyl, H. (1951). A half-century of mathematics. The American Mathematical Monthly 58,
523–553.
17