lecture notes on hilbert spaces and quantum mechanicslandsman/hsqm.pdf · viiquantum mechanics and...

82
1 Lecture Notes on Hilbert Spaces and Quantum Mechanics Draft: December 20, 2004 N.P. Landsman Department of Mathematics Radboud University Nijmegen Postbus 9010 6500 GL NIJMEGEN THE NETHERLANDS email: [email protected] website: http://www.math.kun.nl/landsman/HSQM.html tel.: 024-3652874 office: N2019c

Upload: vanhanh

Post on 15-Aug-2018

227 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

1

Lecture Notes on

Hilbert Spaces and Quantum MechanicsDraft: December 20, 2004

N.P. Landsman

Department of MathematicsRadboud University Nijmegen

Postbus 90106500 GL NIJMEGEN

THE NETHERLANDS

email: [email protected]: http://www.math.kun.nl/∼landsman/HSQM.html

tel.: 024-3652874office: N2019c

Page 2: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

2

Page 3: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

Contents

I Historical notes and overview 5I.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5I.2 Origins in mathematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6I.3 Origins in physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

II Metric spaces, normed spaces, and Hilbert spaces 13II.1 Basic definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13II.2 Convergence and completeness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14II.3 Orthogonality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

IIIOperators and functionals 19III.1 Bounded operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19III.2 Bounded functionals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20III.3 The adjoint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

IV Compact operators 25IV.1 Linear algebra revisited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25IV.2 The spectral theorem for self-adjoint compact operators . . . . . . . . . . . . . . . 26IV.3 Trace-class and Hilbert–Schmidt operators . . . . . . . . . . . . . . . . . . . . . . . 28

V Closed unbounded operators 31V.1 The closure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31V.2 Symmetric and self-adjoint operators . . . . . . . . . . . . . . . . . . . . . . . . . . 33

VI Spectral theory for closed unbounded operators 37VI.1 Resolvent and spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37VI.2 The spectrum of self-adjoint operators . . . . . . . . . . . . . . . . . . . . . . . . . 40

VIIQuantum mechanics and Hilbert space 45VII.1Classical mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45VII.2Quantum mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46VII.3The Schrodinger equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47VII.4Stone’s theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48VII.5States and observables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

VIIISpectral theory for self-adjoint operators 55VIII.1The Stieltjes integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55VIII.2Spectral densities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58VIII.3Operator-valued Stieltjes integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . 61VIII.4The spectral theorem for bounded operators . . . . . . . . . . . . . . . . . . . . . . 63VIII.5The spectral theorem for unbounded operators . . . . . . . . . . . . . . . . . . . . 67VIII.6Self-adjoint extensions of symmetric operators . . . . . . . . . . . . . . . . . . . . . 70

3

Page 4: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

4 CONTENTS

IX Quantum logic 73IX.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73IX.2 Propositional logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73IX.3 The lattice of a propositional logic . . . . . . . . . . . . . . . . . . . . . . . . . . . 75IX.4 The logic of classical physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76IX.5 The logic of quantum mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77IX.6 Gleason’s theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

Page 5: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

Chapter I

Historical notes and overview

I.1 Introduction

The concept of a Hilbert space is seemingly technical and special. For example, the reader hasprobably heard of the space `2 (or, more precisely, `2(Z)) of square-summable sequences of realor complex numbers.1 That is, `2 consists of all infinite sequences . . . , c−2, c−1, c0, c1, c2, . . .,ck ∈ K, for which

∞∑k=−∞

|ck|2 <∞.

Another example of a Hilbert space one might have seen is the space L2(R) of square-integrablecomplex-valued functions on R, that is, of all functions2 f : R → K for which∫ ∞

−∞dx |f(x)|2 <∞.

In view of their special nature, it may therefore come as a surprise that Hilbert spaces playa central role in many areas of mathematics, notably in analysis, but also including (differential)geometry, group theory, stochastics, and even number theory. In addition, the notion of a Hilbertspace provides the mathematical foundation of quantum mechanics. Indeed, the definition of aHilbert space was first given by von Neumann (rather than Hilbert!) in 1927 precisely for thelatter purpose. However, despite his exceptional brilliance, even von Neumann would probablynot have been able to do so without the preparatory work in pure mathematics by Hilbert andothers, which produced numerous constructions (like the ones mentioned above) that are nowregarded as examples of the abstract notion of a Hilbert space. It is quite remarkable how aparticular development within pure mathematics crossed one in theoretical physics in this way;this crossing is reminiscent to the one leading to the calculus around 1670; see below. Today, themost spectacular new application of Hilbert space theory is given by Noncommutative Geometry[5], where the motivation from pure mathematics is merged with the physical input from quantummechanics. Consequently, this is an important field of research in pure mathematics as well as inmathematical physics.

In what follows, we shall separately trace the origins of the concept of a Hilbert space inmathematics and physics. As we shall see, Hilbert space theory is part of functional analysis,an area of mathematics that emerged between approximately 1880–1930. Functional analysis isalmost indistinguishable from what is sometimes called ‘abstract analysis’ or ‘modern analysis,’

1In what follows, we mainly work over the reals in order to serve intuition, but many infinite-dimensional vectorspaces, especially Hilbert spaces, are defined over the complex numbers. Hence we will write our formulae in a waythat is correct also for C instead of R. Of course, for z ∈ R the expression |z|2 is just z2. We will occasionally usethe fancy letter K, for Korper, which in these notes stands for either K = R or K = C.

2As we shall see, the elements of L2(R) are, strictly speaking, not simply functions but equivalence classes ofBorel functions. For detailed information we recommend the course Maat en Integraal by Professor A. van Rooij,which runs parallel to the present one.

5

Page 6: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

6 CHAPTER I. HISTORICAL NOTES AND OVERVIEW

which marked a break with classical analysis. The latter involves, roughly speaking, the studyof properties of a single function, whereas the former deals with spaces of functions.3 One mayargue that classical analysis is tied to classical physics,4 whereas modern analysis is associated withquantum theory. Of course, both kinds of analysis were largely driven by intrinsic mathematicalarguments as well.5 The final establishment of functional analysis and Hilbert space theory around1930 was made possible by combining a concern for rigorous foundations with an interest in physicalapplications [3].

I.2 Origins in mathematics

Cf. [3, 6, 21, 23] for more information on the history of functional analysis and Hilbert spaces. Thekey idea behind functional analysis is to look at functions as points in some infinite-dimensionalvector space. To appreciate the depth of this idea, it should be mentioned that the concept of afinite-dimensional vector space, today routinely taught to first-year students, only emerged in thework of Grassmann between 1844 and 1862 (to be picked up very slowly by other mathematiciansbecause of the obscurity of Grassmann’s writings), and that even the far less precise notion ofa ‘space’ (other than a subset of Rn) was not really known before the work of Riemann around1850. Indeed, Riemann not only conceived the idea of a manifold (albeit in embryonic form, to bemade rigorous only in the 20th century), whose points have a status comparable to points in Rn,but also explicitly talked about spaces of functions (initially analytic ones, later also more generalones). However, Riemann’s spaces of functions were not equipped with the structure of a vectorspace. In 1885 Weierstrass considered the distance between two functions (in the context of thecalculus of variations), and in 1897 Hadamard took the crucial step of connecting the set-theoreticideas of Cantor with the notion of a space of functions. Finally, in his PhD thesis of 1906, whichis often seen as a turning point in the development of functional analysis, Hadamard’s studentFrechet defined what is now called a metric space (i.e., a possibly infinite-dimensional vector spaceequipped with a metric, see below), and gave examples of such spaces whose points are functions.6

After 1914, the notion of a topological space due to Hausdorff led to further progress, eventuallyleading to the concept of a topological vector space, which contains all spaces mentioned below asspecial cases.

To understand the idea of a space of functions, we first reconsider Rn as the space of allfunctions f : 1, 2, . . . , n → R, under the identification x1 = f(1), . . . , xn = f(n). Clearly, underthis identification the vector space operations in Rn just correspond to pointwise operations onfunctions (e.g., f + g is the function defined by (f + g)(k) := f(k) + g(k), etc.). Hence Rn is afunction space itself, consisting of functions defined on a finite set.

The given structure of Rn as a vector space may be enriched by defining the length ‖f‖ of avector f and the associated distance d(f, g) = ‖f − g‖ between two vectors f and g. In addition,

3The modern concept of a function as a map f : [a, b] → R was only arrived at by Dirichlet as late as 1837,following earlier work by notably Euler and Cauchy. But Newton already had an intuitive graps of this concept, atleast for one variable.

4Classical analysis grew out of the calculus of Newton, which in turn had its roots in both geometry and physics.(Some parts of the calculus were later rediscovered by Leibniz.) In the 17th century, geometry was a practicalmatter involving the calculation of lenths, areas, and volumes. This was generalized by Newton into the calculus ofintegrals. Physics, or more precisely mechanics, on the other hand, had to do with velocities and accellerations andthe like. This was abstracted by Newton into differential calculus. These two steps formed one of the most brilliantgeneralizations in the history of mathematics, crowned by Newton’s insight that the operations of integration anddifferentiation are inverse to each other, so that one may speak of a unified differential and integral calculus, orbriefly calculus. Attempts to extend the calculus to more than one variable and to make the ensuing machinerymathematically rigorous in the modern sense of the word led to classical analysis as we know it today. (Newton usedtheorems and proofs as well, but his arguments would be called “heuristic” or “intuitive” in modern mathematics.)

5The jump from classical to modern analysis was as discontinuous as the one from classical to quantum mechanics.The following anecdote may serve to illustrate this. G.H. Hardy was one of the masters of classical analysis and oneof the most famous mathematicians altogether at the beginning of the 20th century. John von Neumann, one of thefounders of modern analysis, once gave a talk on this subject at Cambridge in Hardy’s presence. Hardy’s commentwas: “Obviously a very intelligent man. But was that mathematics?”

6Frechet’s main example was C[a, b], seen as a metric space in the supremum-norm, i.e., d(f, g) = ‖f − g‖ with‖f‖ = supf(x) | x ∈ [a, b].

Page 7: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

I.2. ORIGINS IN MATHEMATICS 7

the angle θ between f and g in Rn is defined. Lengths and angles can both be expressed throughthe usual inner product

(f, g) =n∑k=1

f(k)g(k) (I.1)

through the relations‖f‖ =

√(f, f) (I.2)

and(f, g) = ‖f‖‖g‖ cos θ. (I.3)

In particular, one has a notion of orthogonality of vectors, stating that f and g are orthogonalwhenever (f, g) = 0, and an associated notion of orthogonality of subspaces:7 we say that V ⊂ Rnand W ⊂ Rn are orthogonal if (f, g) = 0 for all f ∈ V and g ∈ W . This, in turn, enables one todefine the (orthogonal) projection of a vector on a subspace of Rn.8 Even the dimension n of Rnmay be recovered from the inner product as the cardinality of an arbitrary orthogonal basis.9

Now replace 1, 2, . . . , n by an infinite set. In this case the corresponding space of functions willobviously be infinite-dimensional in a suitable sense.10 The simplest example is N = 1, 2, . . . , ,so that one may define R∞ as the space of all functions f : N → R, with the associated vectorspace structure given by pointwise operations. However, although R∞ is well defined as a vectorspace, it turns out to be impossible to define an inner product on it, or even a length or distance.Indeed, defining

(f, g) =∞∑k=1

f(k)g(k) (I.4)

it is clear that the associated length ‖f‖ (still given by (I.2)) is infinite for most f . This is hardlysurprising, since there are no growth conditions on f at infinity. The solution is to simply restrictR∞ to those functions with ‖f‖ < ∞. These functions by definition form the set `2(N), which iseasily seen to be a vector space. Moreover, it follows from the Cauchy–Schwarz inequality

(f, g) ≤ ‖f‖‖g‖ (I.5)

that the inner product is finite on `2(N). Consequently, the entire geometric structure of Rn in so faras it relies on the notions of lengths and angles (including orthogonality and orthogonal projections)is available on `2(N). Running ahead of the precise definition, we say that Rn ∼= `2(1, 2, . . . , n)is a finite-dimensional Hilbert space, whereas `2(N) is an infinite-dimensional one. Similarly, onemay define `2(Z) (or indeed `2(S) for any countable set S) as a Hilbert space in the obvious way.

From a modern perspective, `2(N) or `2(Z) are the simplest examples of infinite-dimensionalHilbert spaces, but historically these were not the first to be found.11 The initial motivation forthe concept of a Hilbert space came from the analysis of integral equations12 of the type

f(x) +∫ b

a

dyK(x, y)f(y) = g(x), (I.6)

7A subspace of a vector space is by definition a linear subspace.8This is most easily done by picking a basis ei of the particular subspace V . The projection pf of f onto V is

then given by pf =P

i(ei, f)ei.9This is the same as the cardinality of an arbitrary basis, as any basis can be replaced by an orthogonal one by

the Gram–Schmidt procedure.10The dimension of a vector space is defined as the cardinality of some basis. The notion of a basis is complicated

in general, because one has to distinguish between algebraic (or Hamel) and topological bases. Either way, thedimension of the spaces described below is infinite, though the cardinality of the infinity in question depends onthe type of basis. The notion of an algebraic basis is very rarely used in the context of Hilbert spaces (and moregenerally Banach spaces), since the ensuing dimension is either finite or uncountable. The dimension of the spacesbelow with respect to a topological basis is countably infinite, and for a Hilbert space all possible cardinalities mayoccur as a possible dimension. In that case one may restrict oneself to an orthogonal basis.

11From the point of view of most mathematicians around 1900, a space like `2(N) would have been far to abstractto merit consideration.

12Integral equations were initially seen as reformulations of differential equations. For example, the differentialequation Df = g or f ′(x) = g(x) for unknown f is solved by f =

Rg or f(x) =

R x0 dy g(y) =

R 10 dyK(x, y)g(y) for

K(x, y) = θ(x− y) (where x ≤ 1), which is an integral equation for g.

Page 8: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

8 CHAPTER I. HISTORICAL NOTES AND OVERVIEW

where f , g, and K are continuous functions and f is unknown. Such equations were first studiedfrom a somewhat modern perspective by Volterra and Fredholm around 1900, but the main break-through came from the work of Hilbert between 1904–1910. In particular, Hilbert succeeded inrelating integral equations to an infinite-dimensional generalization of linear algebra by choosingan orthonormal basis ek of continuous functions on [a, b] (such as ek(x) := exp(2πkix) on theinterval [0, 1]), and defining the (generalized) Fourier coefficents of f by fk := (ek, f) with respectto the inner product

(f, g) :=∫ b

a

dx f(x)g(x). (I.7)

The integral equation (I.6) is then transformed into an equation of the type

fk =∑l

Kklfl = gl. (I.8)

Hilbert then noted from the Parseval relation (already well known at the time from Fourier analysisand more general expansions in eigenfunctions)

∑k∈Z

|fk|2 =∫ b

a

dx |f(x)|2 (I.9)

that the left-hand side is finite, so that f ∈ `2(Z). This, then, led him and his students to study `2

also abstractly. E. Schmidt should be mentioned here in particular. Unlike Hilbert, already in 1908he looked at `2 as a ‘space’ in the modern sense, thinking of seqences (ck) as point in this space.Schmidt studied the geometry of `2 as a Hilbert space in the modern sense, that is, empasizingthe inner product, orthogonality, and projections, and decisively contributed to Hilbert’s work onspectral theory.

The space L2(a, b) appeared in 1907 in the work of F. Riesz13 and Fischer as the space of(Lebesgue) integrable functions14 on (a, b) for which∫ b

a

dx |f(x)|2 <∞;

of course, this condition holds if f is continuous on [a, b]. Equipped with the inner product (I.7), thiswas another early example of what is now called a Hilbert space.15 The context of its appearancewas what is now called the Riesz–Fischer theorem: Given any sequence (ck) of real (or complex)numbers and any orthonormal system (ek) in L2(a, b),16 there exists a function f ∈ L2(a, b) forwhich (ek, f) = ck if and only if c ∈ `2, i.e., if

∑k |ck|2 <∞.

At the time, the Riesz–Fischer theorem was completely unexpected, as it proved that two seem-ingly totally different spaces were ‘the same’ from the right point of view. In modern terminology,the theorem establishes an isomorphism of `2 and L2 as Hilbert spaces, but this point of view wasonly established twenty years later, i.e., in 1927, by von Neumann. Inspired by quantum mechan-ics (see below), in that year von Neumann gave the definition of a Hilbert space as an abstractmathematical structure, as follows. First, an inner product on a vector space V over a field K(where K = R or K = C), is a map V × V → K, written as 〈f, g〉 7→ (f, g), satisfying, for allf, g ∈ V and t ∈ K,

1. (f, f) ≥ 0;

2. (g, f) = (f, g);13Frederic Riesz had a brother, Marcel Riesz, who was a well-known mathematician too.14More precisely, the elements of L2(a, b) are not functions but equivalence classes thereof, where f ∼ g when

‖f − g‖2 = 0.15The term ‘Hilbert space’ was first used by Schoenflies in 1908 for `2, and was introduced in the abstract sense

by von Neumann in 1927; see below.16The notion of an orthonormal system of functions on the interval [a, b] was as old as Fourier, and was defined

abstractly by Hilbert in 1906.

Page 9: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

I.2. ORIGINS IN MATHEMATICS 9

3. (f, tg) = t(f, g);

4. (f, g + h) = (f, g) + (f, h);

5. (f, f) = 0 ⇒ f = 0.

Given an inner product on V , one defines an associated length function or norm (see below)‖ · ‖ : V → R+ by (I.2). A Hilbert space (over K) is a vector space (over K) with inner product,with the property that Cauchy sequences with respect to the given norm are convergent (in otherwords, V is complete in the given norm).17 Hilbert spaces are denoted by the letter H rather thanV . Thus Hilbert spaces preserve as much as possible of the geometry of Rn.

It can be shown that the spaces mentioned above are Hilbert spaces. Defining an isomorphismof Hilbert spaces U : H1 → H2 as an invertible linear map preserving the inner product (i.e.,(Uf,Ug)2 = (f, g)1 for all f, g ∈ H1), the Riesz–Fischer theorem shows that `2(Z) and L2(a, b) areindeed isomorphic.

In a Hilbert space the inner product is fundamental, the norm being derived from it. However,one may instead take the norm as a starting point (or, even more generally, the metric, as done byFrechet in 1906). The abstract properties of a norm were first identified by Riesz in 1918 as beingsatisfied by the supremum norm, and were axiomatized by Banach in his thesis in 1922. A normon a vector space V over a field K as above is a function ‖ · ‖ : V → R+ with the properties:

1. ‖f + g‖ ≤ ‖f‖+ ‖g‖ for all f, g ∈ V ;

2. ‖tf‖ = |t|‖f‖; for all f ∈ V and t ∈ K;

3. ‖f‖ = 0 ⇒ f = 0.

The usual norm on Rn satisfies these axioms, but there are many other possibilities, such as

‖f‖p :=

(n∑k=1

|f(k)|p)1/p

(I.10)

for any p ∈ R with 1 ≤ p <∞, or

‖f‖∞ := sup|f(k)|, k = 1, . . . , n.

In the finite-dimensional case, these norms (and indeed all other norms) are all equivalent in thesense that they lead to the same criterion of convergence (technically, they generate the sametopology): if we say that fn → f when ‖fn − f‖ → 0 for some norm on Rn, then this impliesconvergence with respect to any other norm. This is no longer the case in infinite dimension. Forexample, one may define `p(N) as the subspace of R∞ that consists of all vectors f ∈ R∞ for which

‖f‖p :=

( ∞∑k=1

|f(k)|p)1/p

(I.11)

is finite. It can be shown that ‖ · ‖p is indeed a norm on `p(N), and that this space is completein this norm. As with Hilbert spaces, the examples that originally motivated Riesz to give hisdefinition were not `p spaces but the far more general Lp spaces, which he began to study in 1910.For example, Lp(a, b) consists of all (equivalence classes of Lebesgue) integrable functions f on(a, b) for which

‖f‖p :=

(∫ b

a

dx |f(x)|p)1/p

(I.12)

17A sequence (fn) is a Cauchy sequence in V when ‖fn− fm‖ → 0 when n,m→∞; more precisely, for any ε > 0there is N ∈ N such that ‖fn − fm‖ < ε for all n,m > N . A sequence (fn) converges if there is f ∈ V such thatlimn→∞ ‖fn − f‖ = 0.

Page 10: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

10 CHAPTER I. HISTORICAL NOTES AND OVERVIEW

is finite, still for 1 ≤ p <∞, and also ‖f‖∞ := sup|f(x)|, x ∈ (a, b). Eventually, in 1922 Banachdefined what is now called a Banach space as a vector space (over K as before) that is completein some given norm.

Long before the abstract definitions of a Hilbert space and a Banach space were given, peoplebegan to study the infinite-dimensional generalization of functions on Rn. In the hands of Volterra,the calculus of variations originally inspired the study of functions ϕ : V → K, later called func-tionals, and led to early ideas about possible continuity of such functions. However, although thecalculus of variations involved nonlinear functionals as well, only linear functionals turned out tobe tractable at the time (until the emergence of nonlinear functional analysis much later). Indeed,even today (continuous) linear functionals still form the main scalar-valued functions that arestudied on infinite-dimensional (topological) vector spaces. For this reason, throughout this text afunctional will denote a continuous linear functional. For H = L2(a, b), it was independentlyproved by Riesz and Frechet in 1907 that any functional on H is of the form g 7→ (f, g) for somef ∈ H.18 The same result for arbitrary Hilbert spaces H was written down only in 1934–35, againby Riesz, although it is not very difficult.

The second class of ‘functions’ on Hilbert spaces and Banach spaces that could be analyzedin detail were the generalizations of matrices on Rn, that is, linear maps from the given space toitself. Such functions are now called operators.19 For example, the integral equation (I.6) is thensimply of the form (1 + K)f = g, where 1 : L2(a, b) → L2(a, b) is the identity operator 1f = f ,and K : L2(a, b) → L2(a, b) is the operator given by (Kf)(x) =

∫ badyK(x, y)f(y). This is easy

for us to write down, but in fact it took some time before integral of differential equations wereinterpreted in terms of operators acting on functions.20 They managed to generalize practicallyall results of linear algebra to operators, notably the existence of a complete set of eigenvectors foroperators of the stated type with symmetric kernel, that is, K(x, y) = K(y, x).21

The abstract concept of a (bounded) operator (between what we now call Banach spaces) is dueto Riesz in 1913. It turned out that Hilbert and Schmidt had studied a special class of operatorswe now call ‘compact’, whereas an even more famous student of Hilbert’s, Weyl, had investigated asingular class of operators now called ’unbounded’ in the context of ordinary differential equations.Spectral theory and eigenfunctions expansions were studied by Riesz himself for general boundedoperators on Hilbert spaces (seen by him as a special case of general normed spaces), and later,more specifically in the Hilbert space case, by Hellinger and Toeplitz (culminating in their pre-vonNeumann review article of 1927).

In the Hilbert space case, the results of all these authors were generalized almost beyondrecognition by von Neumann in his book from 1932 [19], to whose origins we now turn.

I.3 Origins in physics

For more details on the origin of the Hilbert space concept in physics see [12, 13, 17]. For bio-graphical information on John von Neumann22 see [9, 10, 16, 25].

From 1900 onwards, physicists had begun to recognize that the classical physics of Newton,Maxwell and Lorentz (i.e., classical mechanics, Newtonian gravity, and electrodynamics) could notdescribe all of Nature. The fascinating era that was thus initiated by Planck, to be continued

18More generally, in 1910 Riesz showed that any functional on Lp(a, b) is given by an element Lq(a, b), where1/p+1/q = 1, by the same formula. Since p = 2 implies q = 2, this of course implies the earlier Hilbert space result.

19Or linear operators, but for us linearity is part of the definition of an operator.20For example, Hilbert and Schmidt did not really have the operator concept but (from the modern point of view)

worked in terms of the associated quadratic form. That is, the operator a : H → H defines a map q : H ×H → Kby 〈f, g〉 7→ (f, ag).

21The associated quadratic form then satisfies q(f, g) = q(g, f).22John von Neumann (1903–1957) was a Hungarian prodigy; he wrote his first mathematical paper at the age of

seventeen. Except for this first paper, his early work was in set theory and the foundations of mathematics. In theFall of 1926, he moved to Gottingen to work with Hilbert, the most prominent mathematician of his time. Around1920, Hilbert had initiated his Beweistheory, an approach to the axiomatization of mathematics that was doomedto fail in view of Godel’s later work. However, at the time that von Neumann arrived, Hilbert was mainly interestedin quantum mechanics; see below.

Page 11: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

I.3. ORIGINS IN PHYSICS 11

mainly by Einstein, Bohr, and De Broglie, ended in 1925–1927 with the discovery of quantummechanics. This theory replaced classical mechanics, and was initially discovered in two guises.

First, Heisenberg discovered a form of quantum mechanics that at the time was called ‘matrixmechanics.’ Heisenberg’s basic idea was that in atomic physics physical observables (that is,measurable quantities) should not depend on continuous variables like position and momentum(as he did not believe the concept of an electronic orbit in an atom made sense), but on discretequantities, like the natural numbers n = 1, 2, 3, . . . labelling the orbits in Bohr’s model of the atom.Specifically, Heisenberg thought that in analogy to a ‘quantum jump’ from one orbit to the other,everything should be expressed in terms of two such numbers. Thus he replaced the functionsf(x, p) of position and momentum in terms of which classical physics is formulated by quantitiesf(m,n). In order to secure the law of conservation of energy in his new mechanics, he was forcedto postulate the multiplication rule f ∗ g(m,n) =

∑l f(m, l)g(l, n), replacing the rule fg(x, p) =

f(x, p)g(x, p) of classical mechanics. He noted that f ∗ g 6= g ∗ f , unlike in classical mechanics, andsaw in this non-commutativity of physical observables the key revolutionary character of quantummechanics. When he showed his work to his boss Born, a physicist who as a former assistantto Hilbert was well versed in mathematics, Born saw, after a sleepless night, that Heisenberg’smultiplication rule was the same as the one known for matrices, but now of infinite size.23 ThusHeisenberg’s embryonic formulation of quantum theory, written down in 1925 in a paper generallyseen as the birth of quantum mechanics, came to be known as ‘matrix mechanics’.

Second, Schrodinger was led to a formulation of quantum theory called ‘wave mechanics’, inwhich the famous symbol Ψ, denoting a ‘wave function,’ played an important role. To summarize along story, Schrodinger based his work on de Broglie’s idea that in quantum theory a wave should beassociated to each particle; this turned Einsteins’s concept of a photon from 1905 on its head.24 DeBroglie’s waves should, of course, satisfy some equation, similar to the fundamental wave equationor Maxwell’s equations. It is this equation that Schrodinger proposed in 1926 and which is nownamed after him.25 Schrodinger found his equation by studying the transition from wave optics togeometric optics, and by (wrongly) believing that there should be a similar transition from wavemechanics to classical mechanics.26

Thus in 1926 one had two alternative formulations of quantum mechanics, which looked com-pletely different, but each of which could explain certain atomic phenomena. The relationship andpossible equivalence between these formulations was, of course, much discussed at the time. Themost obvious difficulty in relating Heisenberg’s work to Schrodinger’s was that the former was atheory of observables lacking the concept of a state, whereas the latter had precisely the oppositefeature: Schrodinger’s wave functions were states, but where were the observables? To answer thisquestion, Schrodinger introduced his famous expressions Q = x (more precisely, QΨ(x) = xΨ(x))and P = −i~∂/∂x, defining what we now call unbounded operators on the Hilbert space L2(R3).Subsequently, Dirac, Pauli, and Schrodinger himself recognized that wave mechanics was related tomatrix mechanics in the following way: Heisenberg’s matrix x(m,n) was nothing but the matrixelement (en, Qem) of the position operator Q with respect to the orthonormal basis of L2(R3)given by the eigenfunctions of the Hamiltonian H = P 2/2m+ V (Q). Conversely, the vectors in `2

on which Heisenberg’s matrices acted could be interpreted as states. However, these observationsfell far short of an equivalence proof of wave mechanics and matrix mechanics (as is sometimesclaimed), let alone of a mathematical understanding of quantum mechanics.

Heisenberg’s paper was followed by the ‘Dreimannerarbeit’ of Born, Heisenberg, and Jordan

23At the time, matrices and linear algebra were unknown to practically all physicists.24Einstein’s revolutionary proposal, which marked the true conceptual beginning of quantum theory, had been

that light, universally seen as a wave phenomenon at the time, had a particle nature as well. The idea that lightconsists of particles had earlier been proposed by none other than Newton, but had been discredited after thediscovery of Young around 1800 (and its further elaboration by Fresnel) that light displayes interference phenomenaand therefore should have a wave nature. This was subsequently confirmed by Maxwell’s theory, in which light is anoscillation of the electromagnetic field. In his PhD thesis from 1924, de Broglie generalized and inverted Einstein’sreasoning: where the latter had proposed that light waves are particles, the former postulated that particles arewaves.

25He first found the time-independent Schrodinger equationHΨ = EΨ withH = −~2∆/2m+V , and subsequentlygot the time-dependent one HΨ = i~∂Ψ/∂t.

26Technically, Schrodinger relied on the Hamilton–Jacobi formulation of classical mechanics.

Page 12: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

12 CHAPTER I. HISTORICAL NOTES AND OVERVIEW

(1926); all three were in Gottingen at the time. Born turned to his former teacher Hilbert formathematical advice. Hilbert had been interested in the mathematical structure of physical theoriesfor a long time; his Sixth Problem (1900) called for the mathematical axiomatization of physics.Aided by his assistants Nordheim and von Neumann, Hilbert ran a seminar on the mathematicalstructure of quantum mechanics, and the three wrote a joint paper on the subject (now obsolete).

It was von Neumann alone who, at the age of 23, recognized the mathematical structure ofquantum mechanics. In this process, he defined the abstract concept of a Hilbert space discussedabove; as we have said, previously only some examples of Hilbert spaces had been known. VonNeumann saw that Schrodinger’s wave functions were unit vectors in the Hilbert space L2(R3),and that Heisenberg’s observables were linear operators on the Hilbert space `2. The Riesz–Fischertheorem then implied the mathematical equivalence between wave mechanics and matrix mechan-ics. In a series of papers that appeared between 1927–1929, von Neumann defined Hilbert space,formulated quantum mechanics in this language, and developed the spectral theory of boundedas well as unbounded normal operators on a Hilbert space. This work culminated in his book[19], which to this day remains the definitive account of the mathematical structure of elementaryquantum mechanics.27

Von Neumann proposed the following mathematical formulation of quantum mechanics. Theobservables of a given physical system are the self-adjoint (possibly unbounded) linear operatorsa on a Hilbert space H. The pure states of the system are the unit vectors in H. The expectationvalue of an observable a in a state ψ is given by (ψ, aψ). The transition probability between twostates ψ and ϕ is |(ψ,ϕ)|2. As we see from (I.3), this number is just (cos θ)2, where θ is theangle between the unit vectors ψ and ϕ. Thus the geometry of Hilbert space has a direct physicalinterpretation in quantum mechanics, surely one of von Neumann’s most brilliant insights. Lateron, he would go beyond his Hilbert space approach to quantum theory by developing such topicsand quantum logic (see [1]) and operator algebras (cf. [5]).

27Von Neumann’s book was preceded by Dirac’s The Principles of Quantum Mechanics (1930), which containsanother brilliant, but this time mathematically questionable account of quantum mechanics in terms of linear spacesand operators.

Page 13: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

Chapter II

Metric spaces, normed spaces, andHilbert spaces

II.1 Basic definitions

We repeat two basic definitions from the Introduction, and add a third:

Definition II.1 Let V be a vector space over a field K (where K = R or K = C).An inner product on V is a map V × V → K, written as 〈f, g〉 7→ (f, g), satisfying, for all

f, g, h ∈ V and t ∈ K:

1. (f, f) ∈ R+ := [0,∞) (positivity);

2. (g, f) = (f, g) (symmetry);

3. (f, tg) = t(f, g) (linearity 1);

4. (f, g + h) = (f, g) + (f, h) (linearity 2);

5. (f, f) = 0 ⇒ f = 0 (positive definiteness).

A norm on V is a function ‖ · ‖ : V → R+ satisfying, for all f, g, h ∈ V and t ∈ K:

1. ‖f + g‖ ≤ ‖f‖+ ‖g‖ (triangle inequality);

2. ‖tf‖ = |t|‖f‖ (homogeneity);

3. ‖f‖ = 0 ⇒ f = 0 (positive definiteness).

A metric on V is a function d : V × V → R+ satisfying, for all f, g, h ∈ V :

1. d(f, g) ≤ d(f, h) + d(h, g) (triangle inequality);

2. d(f, g) = d(g, f) for all f, g ∈ V (symmetry);

3. d(f, g) = 0 ⇔ f = g (definiteness).

The notion of a metric applies to any set, not necessarily to a vector space, like an inner productand a norm. These structures are related in the following way:1

1Apart from a norm, an inner product defines another structure called a transition probability, which is ofgreat importance to quantum mechanics; cf. the Introduction. Abstractly, a transition probability on a set S isa function p : S × S → [0, 1] satisfying p(x, y) = 1 ⇔ x = y (cf. Property 3 of a metric) and p(x, y) = p(y, x).See [14]. Now take the set S of all vectors in a complex inner product space that have norm 1, and define anequivalence relation on S by f ∼ g iff f = zg for some z ∈ C with |z| = 1. (Without taking equivalence classesthe first axiom would not be satisfied). The set S = S/ ∼ is then equipped with a transition probability definedby p([f ], [g]) := |(f, g)|2. Here [f ] is the equivalence class of f with ‖f‖ = 1, etc. In quantum mechanics vectorsof norm 1 are (pure) states, so that the transition probability between two states is determined by their angle θ.(Recall the elementary formula from Euclidean geometry (x, y) = ‖x‖‖y‖ cos θ, where θ is the angle between x andy in Rn.)

13

Page 14: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

14 CHAPTER II. METRIC SPACES, NORMED SPACES, AND HILBERT SPACES

Proposition II.2 1. An inner product on V defines a norm on V by means of ‖f‖ =√

(f, f).

2. A norm on V defines a metric on V through d(f, g) := ‖f − g‖.

The proof of this claim is an easy exercise; part 1 is based on the Cauchy–Schwarz inequality

|(f, g)| ≤ ‖f‖‖g‖, (II.1)

whose proof in itself is an exercise, and part 2 is really trivial: the three axioms on a normimmediately imply the corresponding properties of the metric. The question arises when a normcomes from an inner product in the stated way: this question is answered by the Jordan–vonNeumann theorem:

Theorem II.3 A norm ‖·‖ on a vector space comes from an inner product through ‖f‖ =√

(f, f)if and only if

‖f + g‖2 + ‖f − g‖2 = 2(‖f‖2 + ‖g‖2). (II.2)

In that case, one has4(f, g) = ‖f + g‖2 − ‖f − g‖2 for K = R

and4(f, g) = ‖f + g‖2 − ‖f − g‖2 + i‖f − ig‖2 − i‖f + ig‖2 for K = C.

We leave the proof of this theorem as an exercise as well, though it is by no means trivial.Applied to the `p and Lp spaces mentioned in the introduction, this yields the result that the

norm in these spaces comes from an inner product if and only if p = 2; see below for a precisedefinition of Lp(Ω) for Ω ⊆ Rn. There is no (known) counterpart of this result for the transitionfrom a norm to a metric.2 It is very easy to find examples of metrics that do not come from anorm: on any vector space (or indeed any set) V the formula d(f, g) = δfg defines a metric notderived from a norm. Also, if d is any metric on V , then d′ = d/(1 + d) is a metric, too: sincecleary d′(f, g) ≤ 1 for all f, g, this metric can never come from a norm.

II.2 Convergence and completeness

The reason we look at metrics in a Hilbert space course is, apart from general education, thatmany concepts of importance for Hilbert spaces are associated with the metric rather than withthe underlying inner product or norm. The main such concept is convergence:

Definition II.4 Let (xn) := xnn∈N be a sequence in a metric space (V, d). We say that xn → x(i.e., (xn) converges to x ∈ V ) when limn→∞ d(xn, x) = 0, or, more precisely: for any ε > 0 thereis N ∈ N such that d(xn, x) < ε for all n > N .

In a normed space, hence in particular in a space with inner product, this therefore means thatlimn→∞ ‖xn − x‖ = 0.3

A sequence (xn) in (V, d) is called a Cauchy sequence when d(xn, xm) → 0 when n,m→∞;more precisely: for any ε > 0 there is N ∈ N such that d(xn, xm) < ε for all n,m > N . Clearly, aconvergent sequence is Cauchy: from the triangle inequality and symmetry one has

d(xn, xm) ≤ d(xn, x) + d(xm, x).

So for given ε > 0 there is N ∈ N such that d(xn, x) < ε/2, etcetera. However, the conversestatement does not hold in general, as is clear from the example of the metric space (0, 1) withmetric d(x, y) = |x − y|: the sequence xn = 1/n does not converge in (0, 1) (for an exampleinvolving a vector space see the exercises). In this case one can simply extend the given space to[0, 1], in which every Cauchy sequence does converge.

2More generally, a metric (on an arbitrary set) defines a so-called topology on this set, but we leave this to theTopology course by F. Clauwens.

3Such convergence is sometimes called strong convergence, in contrast to weak convergence, which for aninner product space means that limn |(y, xn − x)| = 0 for each y ∈ V .

Page 15: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

II.2. CONVERGENCE AND COMPLETENESS 15

Definition II.5 A metric space (V, d) is called complete when every Cauchy sequence converges.

• A vector space with norm that is complete in the associated metric is called a Banach space.

• A vector space with inner product that is complete in the associated metric is called a Hilbertspace.

Like any good definition, this one too comes with a theorem:

Theorem II.6 For any metric space (V, d) there is a complete metric space (V , d) (unique up toisomorphism) containing (V, d) as a dense subspace4 on which d = d. If V is a vector space, thenso is V . If the metric d comes from a norm, then V carries a norm inducing d (so that V , beingcomplete, is a Banach space). If the norm on V comes from an inner product, then V carries aninner product, which induces the norm just mentioned (so that V is a Hilbert space), and whoserestriction to V is the given inner product.

Since this theorem is well known and basic in analysis, we will not give a complete proof, but willjust sketch the main idea. One defines V as the set of all Cauchy sequences (xn) in V , modulothe equivalence relation (xn) ∼ (yn) when limn d(xn, yn) = 0. (When xn and yn converge inV , this means that they are equivalent when they have the same limit.) The metric d on theset of such equivalence classes [xn] := [(xn)] is defined by d([xn], [yn]) := limn d(xn, yn). (Thislimit exists, since using the triangle inequality one easily shows that |d(xn, yn) − d(xm, ym)| ≤d(xn, xm) + d(yn, ym).) The embedding ι : V → V is given by identifying x ∈ V with the Cauchysequence (xn = x∀n), i.e., ι(x) = [xn = x]. It follows that a Cauchy sequence (xn) in V ⊆ Vconverges to [xn], for

limmd(ι(xm), [xn]) = lim

md([xn = xm], [xn]) = lim

mlimnd(xm, xn) = 0

by definition of a Cauchy sequence. Furthermore, one can show that any Cauchy sequence in Vconverges by approximating its elements by elements of V .

If V is a vector space, the corresponding linear structure on V is given by [xn]+[yn] := [xn+yn]and t[xn] := [txn]. If V has a norm, the corresponding norm on V is given by ‖[xn]‖ := limn ‖xn‖.If V has an inner product, the corresponding inner product on V is given by ([xn], [yn]) :=limn(xn, yn).

A finite-dimensional vector space is complete in any possible norm. In infinite dimensions,completeness generally depends on the norm. For example, take V = C([0, 1]). This space iscomplete in the supremum-norm or sup-norm

‖f‖∞ := supf(x) | x ∈ [0, 1], (II.3)

but is incomplete in the L2-norm‖f‖2 :=

√(f, f) (II.4)

derived from the inner product

(f, g) :=∫ 1

0

dx f(x)g(x). (II.5)

The first claim follows from the theory of uniform convergence, whereas the latter follows from anexplicit counterexample to completeness (exercise!).

These considerations can easily be extended to sets Ω ⊆ Rn.5 First, assume that Ω is compact.The space C(Ω) is complete in the supremum-norm (defined by the obvious analogue of (II.3)),but incomplete in the L2-norm induced from the inner product

(f, g) :=∫

Ω

dnx f(x)g(x). (II.6)

4This means that any point in V is the limit of some convergent sequence in V with respect to the metric d.5Strictly speaking, Ω should be a Borel subset of Rm; just think of open or closed subsets, or countable intersec-

tions thereof.

Page 16: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

16 CHAPTER II. METRIC SPACES, NORMED SPACES, AND HILBERT SPACES

To cover the noncompact case, we introduce the spaces Cc(Ω) and C0(Ω). First, Cc(Ω) consists ofall continuous functions on Ω with compact support, where the support of a function is definedas the smallest closed set outside which it vanishes. In other words, Cc(Ω) contains those functionsthat are nonzero inside some compact set. Second, C0(Ω) consists of all continuous functions onΩ that vanish at infinity, in the sense that for each ε > 0 there is a compact subset K ⊂ Ω suchthat |f(x)| < ε for all x outside K. Clearly, Cc(Ω) ⊆ C0(Ω), with equality Cc(Ω) = C0(Ω) = C(Ω)when Ω is compact. Now, when Ω is noncompact it can be shown (by easy examples) that Cc(Ω)is not complete in the sup-norm; its completion turns out to be C0(Ω).6 Also, the inner product(II.6) is still defined on Cc(Ω) (but not on C0(Ω)!), which space again fails to be complete in theassociated norm.

Definition II.7 Let Ω ⊆ Rn.7 The Hilbert space L2(Ω) is defined as the completion of Cc(Ω) inthe inner product (II.6).

This is a fairly abstract definition of this space, as it relies on the completion procedure explainedin the proof of Theorem II.6.8

II.3 Orthogonality

As stressed in the Introduction, Hilbert spaces are the vector spaces whose geometry is closest tothat of R3. In particular, the inner product yields a notion of orthogonality. We say that twovectors f, g ∈ H are orthogonal, written f ⊥ g, when (f, g) = 0.9 Similary, two subspaces10

K ⊂ H and L ⊂ H are said to be orthogonal (K ⊥ L) when (f, g) = 0 for all f ∈ K and all g ∈ L.A vector f is called orthogonal to a subspace K, written f ⊥ K, when (f, g) = 0 for all g ∈ K,etc. We define the orthogonal complement K⊥ of a subspace K ⊂ H as

K⊥ := f ∈ H | f ⊥ K. (II.7)

This set is automatically linear, so that the map K 7→ K⊥, called orthocomplementation, is anoperation from subspaces of H to subspaces of H. Clearly, H⊥ = 0 and 0⊥ = H.

Lemma II.8 For any subspace K ⊂ H one has K⊥ = K⊥

= K⊥.

If fn → f in H, then (fn − f, g) → 0∀g ∈ H by Cauchy–Schwarz. Let g ∈ K and fn ∈ K⊥; itfollows that (f, g) = limn(fn, g) = 0. Hence f ∈ K⊥. The other equality is trivial.

This lemma will be improved later on.Recall that a basis of a finite-dimensional vector space V is a linearly independent set (ei) such

that any v ∈ V can be written as v =∑i viei for some vi ∈ K. In the infinite-dimensional case,

we restrict ourselves to the case that V is a Banach space, and define a basis in precisely the sameway, where now the sum

∑i viei may be infinite, and is meant to converge to v in norm. In other

words, a basis of a Banach space V has the property that the linear span of all ei is dense inV . We say that V is separable when it contains a countable dense subset; this is easily seen tobe equivalent to having a countable basis. This is a pretty subtle notion: for example, C0(Ω) isseparable in the sup-norm, but Cb(Ω) (i.e., the space of all bounded continuous functions on Ω) isnot.

When V is a Hilbert space, we alway require a basis (ei) to be orthonormal, i.e., (ei, ej) = δij .Using Zorn’s lemma (or the axiom of choice), it can be shown that any Hilbert space H has an

6For those who know some topology: this is true for any locally compact Hausdorff space Ω.7Again, as a Borel subset.8For those familiar with modern integration theory (i.e., in the sense of Lebesgue), a direct definition is available:

L2(Ω) consists of equivalence classes of Borel functions f : Ω → K that satisfyRΩ d

nx |f(x)|2 <∞. The equivalence

relation is f ∼ g whenRΩ d

nx |f(x) − g(x)|2 = 0. The inner product is still given by (II.6), with two comments:Firstly, the integral is defined in the sense of Lebesgue (whereas in the main text we could use the integral in thesense of Riemann, since (II.6) was only defined for continuous functions). Secondly, the left-hand side is really([f ], [g]) instead of (f, g).

9By definition of the norm, if f ⊥ g one has Pythagoras’ theorem ‖f + g‖2 = ‖f‖2 + ‖g‖2.10Recall that a subspace of a vector space is by definition a linear subspace.

Page 17: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

II.3. ORTHOGONALITY 17

orthonormal basis. The cardinality of all bases is the same, and defines the dimension of H.However, in this course we restrict ourselves to separable Hilbert spaces. These are sufficient todeal with quantum mechanics and with most applications of Hilbert space theory to mathematics.For example, the spaces L2(Ω) for (Borel) Ω ⊂ Rn and `2(S) for countable S are separable. In theseparable case, many arguments simplify. For example, one can prove existence of an orthonormalbasis without Zorn’s lemma:

Theorem II.9 Any separable Hilbert space has an orthonormal basis.

The proof is an exercise, based on the Gram–Schmidt procedure.Let (ei) by an orthonormal basis of a separable Hilbert space H. By definition, any f ∈ H can

be written as f =∑i ciei. Taking the inner product with a fixed ej , one has

(ej , f) = (ej , limN→∞

N∑i=1

ciei) = limN→∞

ci(ej , ei) = cj .

Here we have assumed N > k, and the limit may be taken outside the inner product since if fn → fin H then (g, fn) → (g, f) for fixed g ∈ H, as follows from Cauchy–Schwarz. It follows that

f =∑i

(ei, f)ei, (II.8)

from which one obtains Parseval’s equality∑i

|(ei, f)|2 = ‖f‖2. (II.9)

Page 18: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

18 CHAPTER II. METRIC SPACES, NORMED SPACES, AND HILBERT SPACES

Page 19: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

Chapter III

Operators and functionals

III.1 Bounded operators

Let V be a vector space. An operator on V is a linear map a : V → V (i.e., a(λv + µw) =λa(v) + µa(w) for all λ, µ ∈ K and v, w ∈ V ). We usually write av for a(v).

Definition III.1 If V is a normed vector space, we call an operator a : V → V bounded when

‖a‖ := sup ‖av‖, v ∈ V, ‖v‖ = 1 <∞. (III.1)

It easily follows that if a is bounded, then

‖a‖ = inf C ≥ 0 | ‖av‖ ≤ C‖v‖ ∀v ∈ V . (III.2)

Moreover, if a is bounded, then‖av‖ ≤ ‖a‖ ‖v‖ (III.3)

for all v ∈ V . This inequality is extremely important.1 Another useful property, which immediatelyfollows from (III.3), is

‖ab‖ ≤ ‖a‖‖b‖. (III.4)

Here ab := a b, so that (ab)(v) := a(bv).When V is finite-dimensional, any operator is bounded. There are many ways to see this.2

Indeed, let V = Cn, so that B(Cn) = Mn(C) is the space of all complex n × n matrices. TheHilbert space norm on Cn is the usual one, i.e., ‖z‖2 =

∑nk=1 zkzk, and the induced norm on

Mn(C) is (III.1). As we shall see (cf. (III.14) below), ‖a‖2 is equal ‖a∗a‖ (where a∗ is the adjointof a), which also coincides with the largest eigenvalue of a∗a; this is immediate from the minimax-property of eigenvalues.3

For an infinite-dimensional example, take V = Cc(R), and let a ∈ C(R) (with abuse of notation)define the multiplication operator a0 : Cc(R) → Cc(R) by a0f(x) := a(x)f(x). This operatoris bounded with respect to the sup-norm or the L2-norm on Cc(R) iff a ∈ Cb(R), in other words, if‖a‖∞ <∞. For example, a0 is not bounded when a(x) = x, and bounded when a(x) = exp(−x2).Let us now pass to the closure of V in the two norms at hand. The closure of Cc(R) in the sup-normis C0(R), whereas its closure in the L2-norm is L2(R). Call the resulting multiplication operatorsimply a. In either case, one then has ‖a‖ = ‖a‖∞. This generalizes to any Ω ⊆ Rn (exercise forcompact Ω). To prove this and similar results, the following argument is very useful.

We denote the completion or closure of a subspace V0 of a Banach space V by V0 ⊂ V ; theabstract completion of V0 can be identified with a subspace of V0 by identifying the equivalenceclass [xn] in the abstract completion procedure with the limit of (xn) in V (check the details!).

1For those who know topology: it can be shown that an operator on V is bounded iff it is continuous in thetopology on V derived from the norm.

2For topologists: any linear map Kn → Kn is automatically continuous, hence bounded (the continuity of themap on the unit sphere implies that the supremum in (III.1) is finite, since the unit ball in Kn is compact).

3Since a∗a is Hermitian, it has a basis of eigenvectors.

19

Page 20: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

20 CHAPTER III. OPERATORS AND FUNCTIONALS

Proposition III.2 Let V0 ⊂ V be a subspace of a Banach space, and let a0 : V0 → V0 be a boundedoperator. Then there exists a unique bounded extension of a0 to an operator a : V0 → V0 with thesame norm as a0. In particular, when V0 is dense in V , the extension a is defined on all of V .

The idea is to define ax for x /∈ V0 by ax := limn a0xn, where (xn) ⊂ V0 converges to x. Weleave the easy details to the reader.4 Hence in the above examples it suffices to compute the normof a0 in order to find the norm of a.

Proposition III.3 If V is a normed vector space, then so is the set B(V ) of all bounded operatorson V . If V is a Banach space, then so is B(V ).5

The proof of the first claim is trivial, given (III.3). For the second claim, Let an be a Cauchysequence in B(V ). In other words, for any ε > 0 there is a natural number N(ε) such that‖an − am‖ < ε when n,m > N(ε). For arbitrary v ∈ V , the sequence anv is a Cauchy sequencein V , because

‖anv − amv‖ ≤ ‖an − am‖ ‖v‖ ≤ ε‖v‖ (III.5)

for n,m > N(ε). Since V is complete by assumption, the sequence anv converges to some w ∈ V .Now define a map a on V by av := w = limn anv. This map is obviously linear. Taking n→∞ in(III.5), we obtain

‖av − amv‖ ≤ ε‖v‖ (III.6)

for all m > N(ε) and all v ∈ V . It now follows from (III.1) that a − am is bounded. Sincea = (a− am) + am, and B(V ) is a linear space, we infer that a is bounded. Moreover, (III.6) and(III.1) imply that ‖a− am‖ ≤ ε for all m > N(ε), so that an converges to a. Since we have justseen that a ∈ B(V ), this proves that B(V ) is complete.

III.2 Bounded functionals

The second class of functions on a vector space V worthy of study consists of functionals, that is,linear maps ϕ : V → K. The definition of boundedness of functionals defined on a normed vectorspace is basically the same as for operators: we say that ϕ is bounded when

‖ϕ‖ := sup |ϕ(v)|, v ∈ V, ‖v‖ = 1 <∞, (III.7)

and as for operators we have

‖ϕ‖ = inf C ≥ 0‖ | ϕ(v)‖ ≤ C‖v‖ ∀v ∈ V (III.8)

and, if ϕ is bounded,‖ϕ(v)‖ ≤ ‖ϕ‖ ‖v‖ (III.9)

for all v ∈ V .6 Similarly, if V is a Banach space then so is the space V ∗ of all bounded functionalson V ; this space is called the dual of V . Finally, the obvious analogue of Proposition III.2 holds:if V0 ⊂ V is a subspace of a Banach space, and ϕ0 : V0 → K is a bounded functional, then there

4When V is a Hilbert space H, and H0 ⊂ H, then a bounded operator a0 : H0 → H0 cannot merely beisometrically extended to the closure H0, but even, still isometrically, to all of H. This extension is not unique; anexample is given by pap, where p is the projection onto H0. See below.

5Combining this result with (III.4), one arrives at the conclusion that B(V ) is a so-called Banach algebra. Recallthat an algebra is a vector space with an associative bilinear operation (‘multiplication’) · : A×A→ A; we usuallywrite ab for a · b. It is clear that B(V ) is an algebra (with unit) under operator multiplication. A Banach algebrais a Banach space A which is at the same time an algebra, in which for all a, b ∈ A one has (III.4). It follows thatmultiplication in a Banach algebra is separately continuous in each variable.

6As for operators, a bounded functional is automatically continuous. In fact, in the literature the term ‘boundedfunctional’ is rather unusual, to be replaced by ‘continuous functional’.

Page 21: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

III.3. THE ADJOINT 21

exists a unique extension of ϕ0 to a bounded functional ϕ : V0 → V0 with the same norm as ϕ0.Hence when V0 is dense in V , the extension ϕ is defined on V .7

Returning to our example V = Cc(R), a function a ∈ C(R) does not merely define an operatoron V but also a functional ϕa : V → K by the formula ϕa(f) :=

∫R dx a(x)f(x). One can show (see

below) that ϕa is bounded with respect to the sup-norm iff a ∈ L1(R), i.e., iff ‖a‖1 :=∫

R dx |a(x)| <∞, with norm ‖ϕa‖ = ‖a‖1, whereas ϕa is bounded with respect to the L2-norm iff a ∈ L2(R),with norm ‖ϕa‖ = ‖a‖2. This follows from the theorem of Riesz, in which Lp(Ω) for Ω ⊂ Rn isdefined as the completion of Cc(Ω) in the Lp-norm

‖f‖p :=(∫

Ω

dnx |f(x)|p)1/p

(III.10)

for 1 ≤ p <∞, and‖f‖∞ := supf(x) | x ∈ Ω. (III.11)

Theorem III.4 Let 1 ≤ p ≤ ∞. The dual of Lp(Ω) is Lq(Ω), where 1p + 1

q = 1. The identificationof f ∈ Lq(Ω) with an element ϕf of Lp(Ω)∗ is given by ϕf (g) =

∫Ωdnx f(x)g(x).

We omit the proof of this theorem, as it relies on Lebesgue integration theory. We will, however,prove the Hilbert space case p = 2 (hence q = 2). Namely, this is a special case of one of the mostimportant results on Hilbert spaces,8 which also confirms the fact that Hilbert spaces play a uniquerole among general Banach spaces:

Theorem III.5 For any Hilbert space H one has H∗ ∼= H, where for K = C the isomorphismH → H∗ is antilinear.9 In other words, any f ∈ H defines a bounded functional ϕf on H byϕf (g) = (f, g), and any element of H∗ is of this form. Moreover, ‖ϕf‖ = ‖f‖.

The proof is easy. First, given f ∈ H, ϕf is bounded by Cauchy–Schwarz. Conversely, takeϕ ∈ H∗, and let N be the kernel of ϕ. This is a closed subspace of H by the boundedness of ϕ.If N = H then ϕ = 0 so we are ready, since ϕ = ϕf=0. Assume N 6= H. Since, N is closed, N⊥

is not empty, and contains h with ‖h‖ = 1.10 For any f ∈ H, one has ϕ(f)h − ϕ(h)f ∈ N , so(h, ϕ(f)h− ϕ(h)f) = 0, which means ϕ(f)(h, h) = ϕ(h)(h, f), or ϕ(f) = (g, f) with g = ϕ(h)h.

To compute ‖ϕf‖, first use Cauchy–Schwarz to prove ‖ϕf‖ ≤ ‖f‖, and then apply ϕf to f toprove equality.

III.3 The adjoint

Now let V = H be a Hilbert space, and let a : H → H be a bounded operator. The inner producton H gives rise to a map a 7→ a∗, which is familiar from linear algebra: if H = Kn, so that, uponchoosing the standard basis (ei), a is a matrix a = (aij) with aij = (ei, aej), then the adjoint isgiven by a∗ = (aji). In other words, one has

(a∗f, g) = (f, ag) (III.12)

7This result is completely overshadowed by the Hahn–Banach theorem: For a functional ϕ0 on a linearsubspace V0 of a Banach space V there exists a functional ϕ on V such that ϕ = ϕ0 on V0 and ‖ϕ‖ = ‖ϕ0‖. Inother words, each functional defined on a linear subspace of V has an extension to V with the same norm.

Here it is not required that V0 be dense in V , and accordingly the extension ϕ is usually not unique (it is whenV0 is dense, for in that case one defines ϕ for x /∈ V0 by ϕ(x) := limn ϕ0(xn), where (xn) ⊂ V0 converges to x,as for operators). We omit the proof of the Hahn–Banach theorem, for it relies on the axiom of choice and is notparticularly instructive. Note that this result is more general than the corresponding result for operators, as in theHahn–Banach theorem V is not required to be a Hilbert space.

8As mentioned in the Introduction, this theorem is due to Riesz and Frechet for L2-spaces and due to Riesz ingeneral.

9This means that λf 7→ λϕf for λ ∈ C.10Pick an orthonormal basis (en) of N . Since N is closed, any f of the form f =

Pn cnen lies in N , cn ∈ K. Since

N 6= H, there exists g /∈ N , which implies that g 6=P

n(en, g)en, or h := g −P

n(en, g)en 6= 0. Clearly, h ∈ N⊥,and since h 6= 0 the vector h/‖h‖ has norm 1.

Page 22: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

22 CHAPTER III. OPERATORS AND FUNCTIONALS

for all f, g ∈ Kn. This equation defines the adjoint also in the general case, but to prove existenceof a∗ Theorem III.5 is needed: for fixed a ∈ B(H) and f ∈ H, one defines a functional ϕf : H → Kby ϕf (g) := (f, ag). This functional is bounded by Cauchy–Schwarz and (III.3):

|ϕf (g)| = |(f, ag)| ≤ ‖f‖‖ag‖ ≤ ‖f‖‖a‖‖g‖,

so ‖ϕf‖ ≤ ‖f‖‖a‖. Hence by Theorem III.5 there exists a unique h ∈ H such that ϕf (g) = (h, f)for all g ∈ H. Now, for given a the association f 7→ h is clearly linear, so that we may definea∗ : H → H by a∗f := h, or, in other words, by (III.12). Note that the map a 7→ a∗ is anti-linear:for K = C one has (λa)∗ = λa for λ ∈ C. It is an easy exercise to show that for each a ∈ B(H)one has11

‖a∗‖ = ‖a‖ (III.13)‖a∗a‖ = ‖a‖2. (III.14)

A bounded operator a : H → H is called self-adjoint12 when a∗ = a. It immediately followsform (III.12) that for self-adjoint a one has (f, af) ∈ R.13 The most important examples ofself-adjoint operators are projections.

Definition III.6 A projection on a Hilbert space H is an operator p ∈ B(H) satisfying p2 =p∗ = p.

To understand the significance of projections, the reader should first recall the discussion aboutorthogonality and bases in Hilbert spaces in Chapter II.II.3. Now let K ⊂ H be a closed subspaceof H; such a subspace is a Hilbert space by itself, and therefore has an orthonormal basis (ei).Applying (II.8) to K, it is easy to verify that

p : f 7→∑i

(ei, f)ei (III.15)

for each f ∈ H, where the sum converges in H,14 defines a projection. Clearly,

pf = f for f ∈ K;pf = 0 for f ∈ K⊥. (III.16)

Proposition III.7 For each closed subspace K ⊂ H one has H = K⊕K⊥. In other words, givenany closed subspace K ⊂ H each f ∈ H has a unique decomposition f = f‖ + f⊥, where f‖ ∈ Kand f⊥ ∈ K⊥.

The existence of the decomposition is given by f‖ = pf and f⊥ = (1 − p)f , and its uniquenessfollows by assuming f = g‖ + g⊥ with g‖ ∈ K and g⊥ ∈ K⊥: one then has f‖− g‖ = f⊥− g⊥, butsince the left-hand side is in K and the right-hand side is in K⊥, both sides lie in K∩K⊥ = 0.

Conversely, given a projection p, define K := pH. This is a closed subspace of H: if f ∈ pHthen f = pg for some g ∈ H, but then pf = p2g = pg = f , so that f ∈ pH iff pf = f . If fn → ffor fn ∈ pH, then pf = limn pfn = lim fn = f , hence f ∈ pH. Furthermore, K⊥ = (1 − p)H;verifying this fact uses both p∗ = p and p2 = p. Defining f‖ := pf and f⊥ := (1− p)f , one clearlyhas f = f‖ + f⊥ with f‖ ∈ K and f⊥ ∈ K⊥, so this is the unique decomposition of f described inProposition III.7. Hence f‖ is given, for arbitrary f ∈ H, by both the right-hand side of (III.15)and by pf , so that (III.15) holds and p is completely characterized by its image. Hence we haveproved:

11 The second equality turns B(H) into a so-called C∗-algebra. First, An abstract involution on an algebra Ais a real-linear map A→ A∗ such that for all a, b ∈ A and λ ∈ C one has a∗∗ = a, (ab)∗ = b∗a∗, and (λa)∗ = λa∗.A ∗-algebra is an algebra with an involution. A C∗-algebra is a Banach algebra with involution in which (III.14)holds. In addition to B(H), any norm-closed subalgebra A of B(H) that is also closed under the involution (i.e.,if a ∈ A then a∗ ∈ A) is a C∗-algebra. Such algebras are very important throughout physics and mathematics. Infact, it has even been claimed that God is a C∗-algebra.

12Or Hermitian.13In quantum mechanics self-adjoint operators model physical observables, so that these have real expectation

values.14The sum does not converge in the operator norm unless it is finite.

Page 23: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

III.3. THE ADJOINT 23

Theorem III.8 There is a bijective correspondence p↔ K between projections p on H and closedsubspaces K of H: given a projection p one puts K := pH, and given a closed subspace K ⊂ Hone defines p by (III.15), where (ei) is an arbitrary orthonormal basis of K.

An important special case is the unit operator p = 1, associated with K = H.

Corollary III.9 For any subspace K ⊂ H one has K⊥⊥ := (K⊥)⊥ = K. Hence for closedsubspaces K one has K⊥⊥ = K.

First suppose K is closed. By the previous theorem, any f ∈ H, so especially any f ∈ K⊥⊥,has a unique decomposition f = f‖ + f⊥. Since K⊥⊥ = (K⊥)⊥, one must have f⊥ = 0, hencef = f‖ ∈ K. So K⊥⊥ ⊂ K. The converse inclusion is obvious.

If K is not closed, one uses the preceding result and Lemma II.8:

K = K⊥⊥

= (K⊥

)⊥ = K⊥⊥.

Projections are important in many ways. One is their occurrence in the spectral theorem,which will occupy us for the remainder of the course. Furthermore, it follows from PropositionIII.7 that f ∈ H is an eigenvector of p iff f ∈ pH or f ∈ (pH)⊥; in the first case the eigenvectoris 1, and in the second case it is 0. In quantum mechanics a projection p is a so-called yes-noquestion, in the sense that the answer to the question p is yes if the eigenvalue is 1 and no if it is0. For general f with (f, f) = 1, the answer is yes with probability (f, pf) and no with probability(f, (1− p)f) = 1− (f, pf).

Apart from projections, another important class of operators on a Hilbert space consists of theunitaries. An operator u : H → H is called unitary when uu∗ = u∗u = 1. Equivalently, u isisometric, in that (uf, uf) = (f, f) for all f ∈ H, and invertible (with inverse u−1 = u∗). Forexample, if (ei) and (ui) are two orthonormal bases of H, then the operator u(

∑i ciei) :=

∑i ciui is

unitary. In quantum mechanics, one usually encounters unitary operators of the form u = exp(ia),where a is self-adjoint and (for bounded a) the exponential is defined by its usual power seriesexpansion, which converges in operator norm. Clearly, one has u∗ = exp(−ia) and since forcommuting a and b (that is, ab = ba) one has exp(a + b) = exp(a) exp(b), one sees immediatelythat u is indeed unitary.15

A partial isometry is an operator v for which v∗v = p is a projection. A special case is anisometry, characterized by p = 1, i.e., v∗v = 1. An invertible isometry is clearly unitary. Thestructure of partial isometries is as follows.

Proposition III.10 If v is a partial isometry, then v∗ is a partial isometry as well. Let theassociated projection be q := vv∗. The kernel of v is (pH)⊥, and its range is qH. The operator vis unitary from pH to its range qH and zero on (pH)⊥. Conversely, any partial isometry has thisform for projections p and q.

The proof is an exercise.

15In quantum mechanics the operator a is generally unbounded, a case we will deal with in great detail later on.

Page 24: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

24 CHAPTER III. OPERATORS AND FUNCTIONALS

Page 25: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

Chapter IV

Compact operators

IV.1 Linear algebra revisited

Compact operators on a Hilbert space (or, more generally, on a Banach space) are special boundedoperators that behave like matrices on Kn in many ways. To make this point, we first recallthe proof that any hermitian matrix (aij) (i.e., satisfying aji = aij) can be diagonalized. Inlinear algebra this theorem is usually stated in a basis-dependent form. From our more abstractperspective of operators, the matrix (aij) arises from the operator a : Kn → Kn through the choiceof an arbitrary orthonormal basis (ei), in terms of which one has aij = (ei, aej). The spectraltheorem then states that Kn has a (possibly) new basis of eigenvectors (ui), in which a is diagonal:with respect to this basis one has aij = (ui, auj) = (ui, λjuj) = λiδij , where λi are the eigenvaluesof a (possibly degenerate). Now, this result can be restated without any reference to the notion ofa basis, as follows.

Proposition IV.1 Let H = Kn be a finite-dimensional Hilbert space, and let Let a : H → H bea self-adjoint operator on H. There exists a family of mutually orthogonal projections (pα) (i.e.,pαH ⊥ pβH for α 6= β, or pαpβ = δαβ) with

∑α pα = 1 and a =

∑α λαpα, where λα are the

eigenvalues of a. In other words, pα is the projection onto the eigenspace in H with eigenvalue λα;the dimension of pαH is equal to the multiplicity of the eigenvalue λα.

The key to the proof is the following lemma.

Lemma IV.2 Every self-adjoint operator a on H = Kn has an eigenvector with associated eigen-value λ satisfying |λ| = ‖a‖.

Note that by definition of the operator norm an eigenvalue cannot possibly be any bigger!The proof uses some topology, but in the restricted context of H = Kn we simply say (Heine–

Borel) that a set is compact when it is closed and bounded. We will use the following facts:

Lemma IV.3 1. The image of a compact set in H under a continuous map into H or K iscompact.1

2. A continuous function f : K → R on a compact set K attains a maximum and a minimum.

We now prove Lemma IV.2. First, the unit ball B1 ⊂ H, defined in general by

B1 := f ∈ H | ‖f‖ ≤ 1, (IV.1)

is clearly compact in H = Kn. We know from basic analysis that any linear map a : Kn → Kn

is continuous,2 so that aB1 is compact as well. The norm f 7→ ‖f‖ is continuous on H, hence1This true in general: if f : K → Y is a continuous map between a compact space K and an arbitrary topological

space Y , then f(K) is compact in Y .2A linear map V → V on a normed vector space is continuous iff it is continuous at zero. The latter easily follows

for V = Kn. The map is therefore bounded by definition (cf. (III.1)).

25

Page 26: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

26 CHAPTER IV. COMPACT OPERATORS

it follows that the function f 7→ ‖af‖2 attains a maximum on B1, say at f = f0, obviouslywith ‖f0‖ = 1. By definition of the norm (cf. (III.1)), this maximum must be ‖a‖2, so that‖a‖2 = ‖af0‖2. Cauchy–Schwarz and a∗ = a then yield

‖a‖2 = ‖af0‖2 = (af0, af0) = (f0, a2f0) ≤ ‖f0‖‖a2f0‖ ≤ ‖a2‖ = ‖a‖2,

where we have used (III.14). In the Cauchy–Schwarz inequality (I.5) one has equality iff g = zffor some z ∈ K, so that we must have a2f0 = zf0, with |z| = ‖a‖2. Moreover, z ∈ R, as for anyself-adjoint operator eigenvalues must be real (trivial exercise!), so a2f0 = λ2f0 with either λ = ‖a‖or λ = −‖a‖. Now either af0 = λf0, in which case the lemma is proved, or g0 := af0 − λf0 6= 0.In the latter case,

ag0 = a2f0 − λaf0 = λ2f0 − λaf0 = −λg0,

and the lemma follows, too.

Given this lemma, the proof of Proposition IV.1 is peanuts. If (g, f0) = 0 then

(ag, f0) = (g, a∗f0) = (g, af0) = ±λ(g, f0) = 0,

so that a maps f⊥0 into itself. In other words, if p is the projection onto f⊥0 then pa = ap andpa = pap is the restriction of a to f⊥0 = pH. We now iterate the above procedure: we apply exactlythe same reasoning to pa as an operator on pH, finding an eigenvector f1 of pa, which of courseis an eigenvector of a, as a = pa on pH. We then form the orthogonal complement of f1 in pH,etcetera. Since H is finite-dimensional, this procedure ends after n steps, leaving us with a basisf0, . . . , fn−1 of H that entirely consists of eigenvectors by construction. Finally, we assemblethose eigenvectors fk with the same eigenvalue λα and define pα to be the projection onto theirlinear span Kα (i.e., pα is given by (II.8) applied to K = Kα).

IV.2 The spectral theorem for self-adjoint compact opera-tors

Let H be infinite-dimensional (and separable).3 Eigenvectors and eigenvalues of operators a on aHilbert space H are defined in the same way as for H = Kn: if af = λf for some λ ∈ C then f ∈ His called an eigenvector of a with eigenvalue λ. A crucial difference with the finite- dimensionalsituation is that even a bounded self-adjoint operator on an infinite-dimensional Hilbert space maynot have any eigenvectors (let alone a basis of them!). For example, on H = L2(Ω) a multiplicationoperator defined by a nonzero continuous function has no eigenfunctions at all. The idea is nowto define a class of operators on H for which the proof of Proposition IV.1 can be copied, so thatthe existence of a complete set of eigenvectors is guaranteed.

We will once again need some topology, but in our setting of separable Hilbert spaces we maykeeps things simple: a set K in a separable Hilbert space H is compact when every sequence inK has a convergent subsequence, and a map α : H → T , where T = K or T = H, is continuousif it “preserves limits,” i.e., if fn → f in H then α(fn) → α(f) in T .4 For H = Kn, thisnotion of compactness is equivalent to the Heine–Borel property; for separable infinite-dimensionalH this equivalence no longer holds. Our notion of continuity is equivalent to the usual one intopology. The norm f 7→ ‖f‖ is continuous on H, which is tautological given our definition ofcontinuity, because convergence in H has been defined in terms of the norm! I.e., fn → f in Hmeans ‖fn− f‖ → 0, which precisely expresses continuity of the norm. Similarly, according to ourdefinition a bounded operator a : H → H is clearly continuous, too, for ‖afn − af‖ → 0 because‖afn − af‖ ≤ ‖a‖‖fn − f‖ by (III.3). The main point is that Lemma IV.3 still applies.

Definition IV.4 A compact operator on a Hilbert space H is a bounded operator that mapsthe unit ball B1 ⊂ H into a compact set.

3The main theorem below is true in general as well, with a similar proof involving more topology.4Compactness and continuity may be defined for arbitrary topological spaces in this way if one replaces sequences

by nets.

Page 27: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

IV.2. THE SPECTRAL THEOREM FOR SELF-ADJOINT COMPACT OPERATORS 27

This is not the case for any bounded operator, for althoug such operators are continuous, theunit ball in H is not compact.5 If H is finite-dimensional, so that B1 is compact, then any operatoris compact. More generally, a finite-rank operator on an infinite-dimensional Hilbert space (i.e.,an operator whose image is finite-dimensional) is compact.

Proposition IV.5 A bounded operator on a separable Hilbert space is compact iff it is the norm-limit of a sequence of finite-rank operators.

The proof is left as an exercise, as is the following consequence of the proposition.

Corollary IV.6 1. If a is compact then so is its adjoint a∗;

2. If a is compact and b is bounded then ab and ba are compact;

3. A projection p is compact iff it is finite-dimensional.

Using this corollary, it is easy to show that the set of all compact operators in B(H) is aBanach space called B0(H) in the operator norm. Moreover, B0(H) is even an algebra underoperator multiplication, closed under involution.6

Integral operators form an important class of compact operators. Without proof we mentionthat if Ω ⊂ Rn is compact, any operator of the form af(x) =

∫Ωdnx a(x, y)f(y) is compact when

the kernel a(−,−) is continuous on Ω × Ω. More generally, for any Ω ⊂ Rn such an operator iscompact when a(−,−) ∈ L2(Ω × Ω), i.e., when

∫dnxdny |a(x, y)|2 < ∞. The following theorem

completely characterizes self-adjoint compact operators.

Theorem IV.7 Let a be a self-adjoint compact operator on a Hilbert space H. Then H has anorthonormal basis (ei) of eigenvectors of a, in terms of which

a =∑i

λipi, (IV.2)

where pi projects onto the span of ei, and the sum converges in the sense that af =∑i λipif for

each fixed f ∈ H. Moreover, the set (λi) of eigenvalues of a has the property that

limi→∞

|λi| = 0. (IV.3)

Conversely, a bounded self-adjoint operator on H of the form (IV.2) where the eigenvaluessatisfy (IV.3) is compact.

To prove the first claim, we may simply repeat the proof of Proposition IV.1, as compact operatorshave been defined for precisely this purpose! In particular, Lemma IV.2 still holds. However, theinfinite-dimensional situation makes one big difference: the iteration procedure in the proof is nota priori guaranteed to converge. To settle this, let p =

∑i pi (with convergence as understoof in

(II.8)) be the projection onto the subspace of H spanned by all eigenvectors of a. Clearly, a mapspH onto itself, and using a∗ = a, it follows that it maps (pH)⊥ onto itself as well: if g ∈ (pH)⊥

then by definition (f, g) = 0 for all f ∈ pH, so (f, ag) = (a∗f, g) = (af, g) = 0. But (pH)⊥ is aHilbert space in its own right, and since the restriction of a to (pH)⊥ is given by (1 − p)a, thisrestriction is still compact by Corollary IV.6.2. Hence Lemma IV.2 applies to it, implying that(1 − p)a has an eigenvector in (pH)⊥, which is the same as saying that a has an eigenvector in(pH)⊥. This contradicts the definition of pH.

We leave the proof of the second claim as an exercise.

5The unit ball in an infinite-dimensional Hilbert space is not compact in the topology defined by the norm, inwhich fn → f when ‖fn− f‖ → 0. Miraculously, the unit ball is compact in another topology, namely the so-calledweak topology, in which fn → f when |(g, fn − f)| → 0 for each fixed g ∈ H.

6Hence B0(H) is a C∗-algebra; see footnote 11.

Page 28: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

28 CHAPTER IV. COMPACT OPERATORS

Corollary IV.8 Each eigenvalue of a self-adjoint compact operator except possibly 0 has finitemultiplicity, and 0 is the only possibly accumulation point of the set of eigenvalues. Hence we mayrewrite (IV.2) as

a =∑α

λαpα, (IV.4)

where all eigenvalues λα are different and pα is the finite-dimensional projection pα :=∑i|λi=λα

ei.

This expansion has the advantage over (IV.2) that it is unique; in (IV.2) there is an arbitrarinessin the choice of basis within the eigenspace of each eigenvalue with multiplicity greater than one.In particular, the eigenvalues (λi) are uniquely determined by a.

Corollary IV.9 An arbitrary compact operator a on a Hilbert space H has an expansion

af =∑i

µi(ei, f)ui, (IV.5)

where (ei) and (ui) are orthonormal bases of H, µi > 0, and (µ2i ) are the eigenvalues of a∗a.

Since a∗a is self-adjoint, Theorem IV.7 applies, i.e.,

a∗af =∑i

λi(ei, f)ei. (IV.6)

Since a∗a is positive, in the sense that (f, a∗af) ≥ 0 for all f ∈ H (for (f, a∗af) = ‖af‖2), itfollows that λi ≥ 0 for each i. Hence µi =

√λi is well defined as the positive square-root. Now

define ui := µ−1i aei. Using (II.8), eq. (IV.5) is immediate.

The µi are sometimes called the singular values of a.

IV.3 Trace-class and Hilbert–Schmidt operators

Trace-class operators are special compact operators, which are of fundamental importance forquantum mechanics (see Chapter VII below). More generally, for 1 ≤ p < ∞ one defines theSchatten class Bp(H) as the subset of B0(H) consisting of all compact operator for which

‖a‖p :=

(∑i

µpi

)1/p

<∞. (IV.7)

Here the µi are the singular values of a, as just explained. As the notation suggests, ‖ · ‖p is anorm on the linear space Bp(H), which is even complete in this norm (and hence is a Banach spacein its own right. Note that (except when H is finite-dimensional) Bp(H) is not complete in theoperator norm ‖ · ‖; in fact the completion of Bp(H) in ‖ · ‖ is precisely B0(H). It is an exerciseto prove the fundamental inequality

‖a‖ ≤ ‖a‖p. (IV.8)

This is initially defined for for a ∈ B0(H), but if one defines ‖a‖p = ∞ for a ∈ B(H) whenevera /∈ B0(H), it is valid for any bounded operator. Each Bp(H) is an ideal in both B0(H) and inB(H), since, as is easily shown, ab ∈ Bp(H) if a ∈ Bp(H) and b ∈ B(H), or vice versa.

The Schatten classes play an important role in noncommutative geometry for any p [5], butfor us the most important special cases are p = 1 and p = 2. An element of B1(H) is called atrace-class operator, and an element of B2(H) is called a Hilbert–Schmidt operator. Fora ∈ B1(H) we define the trace of a by

Tr a :=∑i

(ei, aei), (IV.9)

Page 29: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

IV.3. TRACE-CLASS AND HILBERT–SCHMIDT OPERATORS 29

where (ei) is an arbitrary orthonormal basis of H. It is easy to show that (IV.9) is absolutelyconvergent and independent of the basis (exercise). Indeed, defining |a| for a ∈ B0(H) by

|a|f :=∑i

µi(ei, f)ei, (IV.10)

where (ei) and (λi) are defined by (IV.6) and µi :=√λi, as before, one has

‖a‖1 = Tr|a|. (IV.11)

In fact, the so-called absolute value |a| of a is defined for any bounded operator (see exercises),and even for any closed unbounded operator. In any case, for a ∈ B1(H) and b ∈ B(H), or viceversa one has

Tr ab = Tr ba, (IV.12)

which impliesTruau∗ = Tr a (IV.13)

for a ∈ B1(H) and u unitary. For a /∈ B1(H), the trace is not defined, not even when it so happensthat

∑i(ei, aei) < ∞ in some orthonormal basis; in that case, the sum will not be absolutely

convergent, and might be infinite in most other bases. In particular, trace-class operators arenecessarily compact and therefore have discrete spectrum (cf. Chapter VI).

Similarly, if a ∈ B2(H), one has

‖a‖22 = Tr a∗a =∑i

‖aei‖2, (IV.14)

for any orthonormal basis (ei). In fact, B2(H) is a Hilbert space in the inner product

(a, b) := Tr a∗b; (IV.15)

this expression is finite by positivity (i.e., Tr a∗a ≥ 0) and Cauchy–Schwarz. The associated normis precisely ‖ · ‖2.

The significance of B1(H) in part comes from the following fundamental result. Following theconvention in quantum mechanics, we will generically denote trace-class operators by ρ.

Theorem IV.10 Identifying ρ ∈ B1(H) with a functional on B0(H) by ρ : a 7→ Tr ρa, one has

B0(H)∗ ∼= B1(H). (IV.16)

Identifying a ∈ B(H) with a functional on B1(H) by a : ρ 7→ Tr ρa, one has

B1(H)∗ ∼= B(H). (IV.17)

Both isomorphisms are isometric bijections between Banach spaces.

Since the proof uses several results we have not discussed, we merely sketch it (see [20] for acomplete proof). The crucial inequality is

|Tr ρa| ≤ ‖ρ‖1‖a‖, (IV.18)

where ρ ∈ B1(H) and a ∈ B(H). This immediately implies B1(H) ⊆ B0(H)∗ and B(H) ⊆ B1(H)∗,under the given identifications. Furthermore, one has

‖a‖2 ≤ ‖a‖1 (IV.19)

and hence B1(H) ⊂ B2(H).To prove the inclusion B0(H)∗ ⊂ B1(H), we fix ω ∈ B0(H)∗. By definition, one has |ω(a)‖ ≤

C‖a‖. Picking a ∈ B2(H) and using (IV.8) for p = 2, this yields |ω(a)‖ ≤ C‖a‖2. Since B2(H) isa Hilbert space, by the Riesz–Frechet theorem one has ω(a) = (b, a) = Tr ba for some b ∈ B2(H).

Page 30: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

30 CHAPTER IV. COMPACT OPERATORS

Further analysis, using the fact that ω must have an extension from B2(H) to B0(H), refines thisto b ∈ B1(H). Then put ρ = b∗.

To prove the inclusion B1(H)∗ ⊂ B(H), we fix ω ∈ B1(H)∗. By definition, one has |ω(ρ)‖ ≤C‖ρ‖1. We now choose ρ to be of the special form ρ = |f〉〈g|, where f, g ∈ H and |f〉〈g|h := (g, h)f .Such operators are trace-class with ‖|f〉〈g|‖1 = ‖f‖‖g‖. Defining a quadratic form q on H (i.e.,a sesquilinear map q : H × H → C) by q(f, g) := ω(|f〉〈g|), we have |q(f, g)| ≤ C‖f‖‖g‖. Suchquadratic forms are said to be bounded, and the fundamental theorem on bounded quadratic formsstates that there is a unique bounded operator a such that q(f, g) = (g, af). But for any a ∈ B(H)and f, g ∈ H we have (g, af) = Tr |f〉〈g|a; to see this, evaluate the trace in an orthonormal basiscontaining a vector proportional to g. Hence ω(ρ) = Tr ρa for ρ of the special form ρ = |f〉〈g|. Butthe span of such operators is dense in B1(H), so that ω has the stated form for any ρ ∈ B1(H).

Page 31: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

Chapter V

Closed unbounded operators

V.1 The closure

We are now familiar with two classes of operators on a Hilbert space: bounded ones, and compactones. As we have seen, the latter are precisely those bounded operators for which a spectral theoryanalogous to that for matrices exists. It would be possible to develop the spectral theory of generalbounded (self-adjoint)operators at this point, but in fact it turns out that this theory can bewritten down almost without extra effort for the far wider class of closed (self-adjoint) operators.Such operators need not be bounded, and “It is a fact of life that many of the most importantoperators which occur in mathematical physics are not bounded” [22].

To come to grips with the notion of un unbounded operator, let us use the definition of continuityin the preceding chapter to prove:

Theorem V.1 An operator on a Hilbert space is bounded iff it is continuous.

Let a : H → H be bounded. By (III.3), if fn → f in H, that is, ‖fn − f‖ → 0, then afn → af inH, since ‖afn − af‖ ≤ ‖a‖‖fn − f‖. Conversely, if a is not bounded, then for each n ∈ N thereis fn ∈ H with ‖fn‖ = 1 and ‖afn‖ ≥ n. The sequence (gn = fn/n) converges to 0, but since‖agn‖ ≥ 1, the sequence (agn) does not converge to a0 = 0. Hence a is not continuous, and theimplication “continuous ⇒ bounded” has been proved by reductio ad absurdum.

It follows that unbounded operators are discontinuous: if fn → f it is not guaranteed thatafn → af ; indeed, it is not guaranteed that (afn) converges at all! Hence an essential differencebetween the bounded and the unbounded case is that whereas for bounded operators a PropositionIII.2 states that even if a is initially defined on some dense subspace D of H, it can uniquely beextended to H by continuity, for unbounded operators a such an extension by continuity will not ingeneral exist. Although a discontinuous extension to H might exist, in practice it is very unusual totalk about unbounded operators a defined on H. This is because interesting unbounded operatorstend to be defined on D ⊂ H by some natural expression, which simply does not make sense on allof H. Consequently, the specification of the subspace D on which a is defined, called the domainof a, denoted by D(a) is absolutely essential when a is an unbounded operator. Unless explicitlystated otherwise, we always assume that the domain of an operator is dense in H (otherwise, wereplace H by the closure of the domain in question).

For example, we have seen that a ∈ C(R) defines a bounded operator on H = L2(R) bymultiplication when ‖f‖∞ <∞, i.e., when a is bounded. When a is unbounded as a function, wecannot really make the natural claim that consequently a is unbounded as a multiplication operatoron H, since a does not map H to itself as soon as it is unbounded. What we can say, however, isthat a is unbounded as a multiplication operator on the subspace D(a) = Cc(R), which (by ourdefinition of L2) is dense in H. Other examples are given by differential operators: whereas onecannnot really say that a = id/dx is unbounded on L2(R) since the derivative of most functions inL2 is not even defined,1 one can say that a is unbounded on, say, D(a) = C1

c (R) := Cc(R)∩C1(R).1Unless one is familiar with the theory of distributions.

31

Page 32: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

32 CHAPTER V. CLOSED UNBOUNDED OPERATORS

There is some flexibility in the choice of the domain, and the theory of unbounded operatorsof closable type even largely revolves around the business of enlarging a given domain. Herewe define an extension of a given operator a : D(a) → H as an operator a1 : D(a1) → Hwhere D(a) ⊂ D(a1) and a1 = a on D(a); we write a ⊂ a1. As we have seen, if a is boundedthere is only one interesting extension, namely the one defined on any larger domain (includingH) by continuity, but in the unbounded case the theory of extension of operators turns out tobe very rich indeed. As a rather trivial example, we could have defined a = id/dx initiallyon D(a) = C∞

c (R) := Cc(R) ∩ C∞(R); the same expression a1 = id/dx but now defined onD(a1) = C1

c (R) is formally an extension of a. Similarly, an unbounded mutliplication operatora ∈ C(R) may initially be defined on D(a) = C∞

c (R), to be extended by the same expressiondefined on D(a1) = Cc(R), as above. Clearly, the largest possible domain on which a multiplicationoperator a ∈ C(Ω) can be defined in the natural way is2

D(a) = f ∈ L2(Ω) | af ∈ L2(Ω). (V.1)

The particular unbounded operators that one can deal with satisfy a relic of continuity. Toexplain this, we define the graph of a as the subset f, af ⊂ H × H. To define the notionof closedness in H × H we equip this set with the structure of a Hilbert space in the followingway: we first equip H × H with the structure of a vector space by defining 〈f1, f2〉 + 〈g1, g2〉 :=〈f1 + g1, f2 + g2〉 and λ〈f1, f2〉 := 〈λf1, λf2〉, and subsequently putting an inner product on it by(〈f1, f2〉, 〈g1, g2〉) := (f1, g1)+(f2, g2). With this structure, H×H is a Hilbert space called H⊕H,for it is precisely the direct sum of H with itself.

Definition V.2 A closed operator a : D(a) → H is a linear map from a dense subspace D(a) ⊂H to H for which either of the following equivalent conditions holds:

• If fn → f in H for a sequence (fn) in D(a) and (afn) converges in H, then f ∈ D(a) andafn → af in H.

• The graph of a is closed.

• The domain D(a) is closed in the norm ‖f‖2a := ‖f‖2 + ‖af‖2.

Note that the norm ‖ ·‖a comes from the new inner product (f, g)a := (f, g)+(af, ag) on D(a).Hence D(a) is a Hilbert space in the new inner product when a is closed.

It is quite unusual for an operator to be closed in the form initially given. For example, amultiplication operator a in L2(Ω) is not closed on C∞

c (Ω) or even on Cc(Ω); as proved below, aturns out to be closed only on the domein (V.1). Fortunately, an operator a that is not closed mayoften be extended into a closed one. The condition for this to be possible is as follows.

Definition V.3 A closable operator a : D(a) → H is a linear map from a dense subspaceD(a) ⊂ H to H with the property that the closure G(a)− of its graph is itself the graph G(a−) ofsome operator a− (called the closure of a).

In other words, a− is a closed extension of a. It is clear that a− is uniquely defined by its graphG(a−) = G(a)−.

Proposition V.4 An operator a is closable iff either of the following equivalent conditions holds:

• If fn → 0 for a sequence (fn) in D(a) and if (afn) converges, then (afn) must converge to 0.

• The closure G(a)− of the graph of a does not contain any element of the form (0, g) for g 6= 0.

In that case, the domain of the closure a− of a is the set of all f ∈ H to which some sequence (fn)in D(a) converges and for which (afn) converges as well. Its action is given by a−f := limn afn.

Finally, the operator a− is the smallest closed extension of a.

2Similarly, the largest possible domain on which ∂/∂xi can be defined in the natural way is D(∂/∂xi) = f ∈L2(Ω) | ∂f/∂xi ∈ L2(Ω), where the derivative is defined in the distributional sense.

Page 33: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

V.2. SYMMETRIC AND SELF-ADJOINT OPERATORS 33

To verify that a− is indeed closed, suppose fn → f and afn → g, with (fn) in D(a−). Sincefn ∈ D(a−) for fixed n, there exists (fm,n) in D(a) such that limm fm,n = fn and limm afm,n =: gnexists. Then clearly limm,n fm,n = f , and we claim that

limm,n

afm,n = g. (V.2)

Namely, ‖afm,n−g‖ ≤ ‖afm,n−afn‖+‖afn−g‖. For ε > 0 take n so that the second term is < ε/2.For that fixed n, a(fm,n−fn) converges as m→∞ because afm,n → gn and afn is independent ofm. Also, recall that fm,n − fn → 0 as m → ∞. By assumption, a is closable, hence by definitionone must have a(fm,n − fn) → 0 in m. Hence we may find m so that ‖afm,n − afn‖ < ε/2, sothat ‖afm,n − g‖ < ε, and (V.2) follows. Hence f ∈ D(a−). Finally, since a−f := limm,n afm,none has a−f = g by (V.2), or a−f = limn afn by definition of g. It follows that a− is closed. Thisextension of a is clearly the minimal closed one.

This argument shows that for a− to be closed it is sufficient that a is closable. Conversely, if afails to be closable it cannot have any closed extension whatsoever, since a− is by definition linear,and a−0 = 0 for any operator. The second condition for closability is then clearly equivalent tothe first.

For example, a multiplication operator a in H = L2(Ω) is closable on both D(a) = C∞c (Ω)

and Cc(Ω): in both cases the closure has domain (V.1). On the other hand, an example of anon-closable operator is given by H = L2(Ω), D(a) = Cc(Ω), and af := f(0)g, where g ∈ H isarbitrary (cf. the exercises).

In general, even closed unbounded operators may have closed extensions, a phenomenon whichis particular important in connection with the theory of self-adjointness; see below.

This section would not be complete without mentioning a special case of one of the most famoustheorems in functional analysis, namely the closed graph theorem (for Hilbert spaces):

Theorem V.5 If a : H → H has a closed graph, then it is bounded. In other words, a closedoperator defined on all of H is necessarily bounded.

We will use this theorem only once, since in applications an operator defined on H is usuallyalready known to be bounded, whereas unbounded operators are never defined on H, so that theirgraph cannot be studied in any case.3

V.2 Symmetric and self-adjoint operators

The most important closed operators, not just for quantum mechanics also but for mathematicsas a whole, are the self-adjoint ones. To define these, we first recall the definition of the adjointa∗ of a bounded operator a : H → H. If we initially define the domain D(a∗) of a∗ as the set ofall f ∈ H for which the functional g 7→ ϕf (g) := (f, ag) is bounded, then the boundedness of aimplies that D(a∗) = H, and we may define a∗ by (III.12), i.e., (a∗f, g) = (f, ag). Requiring thisequality for all g ∈ H defines a∗f uniquely: for we may take g = ei to be a basis vector, so thata∗f =

∑i(f, aei)ei by (II.8). This argument show that the vector a∗f (and hence the operator a∗,

since f ∈ H is arbitrary) is even uniquely defined by (III.12) when g runs over a dense subspacein H.4

Definition V.6 The adjoint a∗ of an unbounded operator a : D(a) → H has domain D(a∗)consisting of all f ∈ H for which the functional g 7→ ϕf (g) := (f, ag) is bounded. On this domain,a∗ is defined by requiring (a∗f, g) = (f, ag) for all g ∈ D(a).

3On the theoretical side, however, an important consequence of the closed graph theorem is another famousresult known as the open mapping theorem: a bounded surjective operator on a Hilbert space H is open (in thatthe image of an open set in H is open; the boundedness of H implies the opposite, namely that the inverse imageof an open set in H is open).

4This can be seen in many ways. For example, any dense subspace D contains a basis of H, as can be proved inthe separable case by the Gram–Schmidt procedure (and in general follows from Zorn’s Lemma). Alternatively, evenif ei /∈ D, the coefficient (h, ei) can be computed by approximating ei by elements in D, since by Cauchy–Schwarz

one has (h, ei) = limn(h, f(i)n ) for any sequence f

(i)n → ei in D.

Page 34: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

34 CHAPTER V. CLOSED UNBOUNDED OPERATORS

Here the vector a∗f exists by the Riesz–Frechet theorem III.5, and it is uniquely determinedby our standing assumption that D(a) be dense in H. However, this time we cannot concludethat D(a∗) = H, as in the bounded case. Indeed, it may even happen that D(a∗) is zero! (Seeexercises.) However, this is pretty pathological, and in most ‘natural’ examples D(a∗) turns out tobe dense. For example, a multiplication operator a ∈ C(Ω) on H = L2(Ω), defined on any of thedomains we have considered (i.e., D(a) = C∞

c (Ω), D(a) = Cc(Ω), or (V.1)), has a∗ = a (i.e., thecomplex conjugate of a seen as a multiplication operator) defined on D(a∗) given by the right-handside of (V.1). The following important result shows when D(a∗) is dense.

Proposition V.7 Let a : D(a) → H (where D(a) is dense) be an operator on a Hilbert space H.

1. The adjoint a∗ is closed.

2. The operator a is closable iff D(a∗) is dense.

3. In that case, one has a− = a∗∗ and (a−)∗ = a∗.

The proof can be elegantly given in terms of the graph G(a). Defining U : H ⊕H → H ⊕H byU(f, g) := (g,−f), it is easy to verify that (exercise!)

G(a∗) = U(G(a)⊥) = U(G(a))⊥. (V.3)

Hence the first claim immediately follows, since any orthogonal complement is closed (cf. LemmaII.8).

By Corollary (III.9) and (V.3) one has G(a)− = G(a)⊥⊥ = (U−1G(a∗))⊥. But

(U−1G(a∗))⊥ = 〈f, g〉 ∈ H ⊕H | (f, a∗h)− (g, h) = 0 ∀h ∈ D(a∗).

Consequently, referring to Proposition V.4, (0, g) ∈ G(a)− iff g ∈ D(a∗)⊥, proving the secondclaim. Using (V.3) twice (first for a∗, then for a), we have

G(a∗∗) = (UG(a∗))⊥ = (U2(G(a)⊥))⊥ = G(a)⊥⊥ = G(a)− = G(a−),

where we have used U2 = −1 and −G(b) = G(b) for any operator b. This proves the first part ofthe third claim. Finally, if a is closable, then using (V.3) and Lemma II.8 and (V.3) once more,we have

G(a∗) = U(G(a)⊥) = U((G(a)−)⊥) = U(G(a−)⊥) = G((a−)∗).

We now come to the central definition of this chapter. An equality a = b between unboundedoperators always stand for D(a) = D(b) and a = b. Similarly, a ⊂ b means D(a) ⊂ D(b) and b = aon D(a).

Definition V.8 Let a : D(a) → H (where D(a) is dense) be an operator on a Hilbert space H.

• If a = a∗, i.e., if D(a∗) = D(a) and (af, g) = (f, ag) for all f, g ∈ D(a), then a is calledself-adjoint.

• If a∗∗ = a∗ (equivalently, if a is closable and a− = a∗ or (a−)∗ = a∗), then a is calledessentially self-adjoint.

• If a ⊂ a∗ i.e., if (af, g) = (f, ag) for all f, g ∈ D(a), then a is called symmetric.

It follows Proposition V.7 that a self-adjoint operator is closed, and that a symmetric operatoris closable (because D(a∗), containing D(a), is dense). For a symmetric operator one has a ⊆a− = a∗∗ ⊆ a∗, with equality at the first position when a is closed, and equality at the secondposition when a is essentially self-adjoint; when both equalities hold, a is self-adjoint. Conversely,an essentially self-adjoint operator is, of course, symmetric. A symmetric operator may or maynot have self-adjoint extensions; we will deal with this problem in detail later on. Without proofwe quote the Hellinger–Toeplitz theorem: If a is self-adjoint on D(a) = H, then a is bounded.This confirms the idea that it is pointless to try to define unbounded operators on all of H in somemanifestly discontinuous way. All this is very subtle, but the following example illustrates at leastthe easy part of the theory:

Page 35: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

V.2. SYMMETRIC AND SELF-ADJOINT OPERATORS 35

Proposition V.9 A real-valued multiplication operator a ∈ C(Ω) on H = L2(Ω) is essentiallyself-adjoint on D(a) = C∞

c (Ω) and on D(a) = Cc(Ω), and is self-adjoint on D(a) = f ∈ L2(Ω) |af ∈ L2(Ω).

Cf. (V.1)). Of course, D(a) = H when ‖a‖∞ <∞, since in that case a is a bounded operator, aswe have seen.5 The proof is almost trivial: by definition, D(a∗) consists of all f ∈ L2(Ω) for wichthe functional g 7→ (f, ag) =

∫Ωdnx f(x)a(x)g(x) is bounded; for each of the three choices of D(a)

listed this implies fa ∈ L2(Ω) by Riesz–Frechet, so that D(a∗) = f ∈ L2(Ω) | af ∈ L2(Ω), witha∗ = a as a multiplication operator. If a is real-valued, then a∗ is self-adjoint by the very sameargument, which implies all claims.

A more difficult example is provided by H = L2([0, 1]) and a = −id/dx; see Section VIII.2 in[22] for the detailed analysis leading to the claims now made. Partial integration shows that a issymmetric on the domain D(a) = f ∈ C1([0, 1]) | f(0) = f(1) = 0, which is easy. It is moredifficult to prove that the adjoint of a with respect to this domain is a∗ = −id/dx defined on thedomain

D(a∗) = f ∈ AC([0, 1]) | df/dx ∈ L2([0, 1]). (V.4)

Here AC([0, 1]) (where AC stands for ‘absolutely continuous’) consists of all functions of the typef(x) =

∫ x0g(x) with g ∈ L1([0, 1]). It turns out that a− is given by a− = −id/dx on D(a−) =

f ∈ D(a∗) | f(0) = f(1) = 0 (hence a is not closed). Since D(a−) 6= D(a∗), a is not essentiallyself-adjoint. It turns out that a has a family (aθ)θ∈[0,2π) of self-adjoint extensions, defined byaθ = −id/dx on the domain D(aθ) = f ∈ D(a∗) | f(1) = exp(iθ)f(0).

5This result may be extended to multiplication operators defined by Lebesgue-measurable functions a : Ω → R.Defining the essential supremum of such a function by ‖a‖ess

∞ : infC | |a(x)| ≤ C for almost everyx ∈ Ω, onehas a ∈ B(L2(Ω)) iff ‖a‖ess

∞ <∞, in that case with ‖a‖ = ‖a‖ess∞ , and Proposition V.9 being valid as it stands.

Page 36: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

36 CHAPTER V. CLOSED UNBOUNDED OPERATORS

Page 37: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

Chapter VI

Spectral theory for closedunbounded operators

Until further notice, we now assume that K = C. We denote the kernel or null space of a map aby N(a) and its range or image by R(a). As before, D(a) denotes the domain of a. Also, a− z forz ∈ C denotes the operator a− z1.

VI.1 Resolvent and spectrum

The theory of the spectrum of a closed operator on a Hilbert space (which may be bounded orunbounded) is a generalization of the theory of eigenvalues of a matrix. From linear algebra werecall:

Proposition VI.1 Let a : Cn → Cn be a linear map. The a is injective iff it is surjective.

This follows from the fundamental fact that if a : V → W is a lnear map between vector spaces,one has R(a) ∼= V/N(a). If V = W = Cn, one one count dimensions to infer that dim(R(a)) =n − dim(N(a)). Surjectivity of a yields dim(R(a)) = n, hence dim(N(a)) = 0, hence N(a) = 0,and vice versa.1

Corollary VI.2 Let a : Cn → Cn be a linear map. Then a − z is invertible (i.e., injective andsurjective) iff z is not an eigenvalue of a, i.e., if there exists no f ∈ Cn such that af = zf .

Defining the spectrum σ(a) of a : Cn → Cn as the set of eigenvalues of a and the resolventρ(a) as the set of all z ∈ C for which a− z is invertible, we therefore have

σ(a) = C\ρ(a). (VI.1)

If z ∈ ρ(a), the equation (a − z)f = g for the unknown f ∈ Cn has a unique solution for any g;existence follows from the surjectivity of a− z, whereas uniqueness follows from its injectivity (ifa− z fails to be injective then any element of its kernel can be added to a given solution).

Now, if a is an operator on an infinite-dimensional Hilbert space, it may not have any eigen-values, even when it is bounded and self-adjoint. For example, if a(x) = exp(−x2) the associatedmultiplication operator a : L2(R) → L2(R) is bounded and self-adjoint, but it has no eigenvalues atall: the equation af = λf for eigenvectors is exp(−x2)f(x) = λf(x) for (almost) all x ∈ R, whichholds only if f is nonzero at a single point. But in that case f = 0 as an element of L2. However,

1In general, this proposition yields the very simplest case of the Atiyah–Singer index theorem, for which theseauthors received the Abel Prize in 2004. We define the index of a linear map a : V → W as index(a) :=dim(ker(a))− dim(coker(a)), where ker(a) = N(a) and coker(a) := W/R(a), provided both quantities are finite. IfV and W are finite-dimensional, Proposition VI.1 yields index(a) = dim(V ) − dim(W ); in particular, if V = Wthen index(a) = 0 for any linear map a. In general, the index theorem expresses the index of an operator in termsof topological data; in this simple case the only such data are the dimensions of V and W .

37

Page 38: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

38 CHAPTER VI. SPECTRAL THEORY FOR CLOSED UNBOUNDED OPERATORS

the situation is not hopeless. More generally, let any a ∈ Cb(R), interpreted as a multiplicationoperator a : L2(R) → L2(R). If x0 ∈ R one may find approximate eigenvectors of a in the followingsense: take

fn(x) := (n/π)1/4e−n(x−x0)2/2. (VI.2)

Then ‖fn‖ = 1 and limn→∞(a(x) − a(x0))fn = 0, although the sequence fn itself has no limit inL2(R). Thus we may call λ = a(x0) something like a generalized eigenvalue of a for any x0 ∈ R,and define the spectrum accordingly: let a : D(a) → H be a (possibly unbounded) operator on aHilbert space. We say that λ ∈ σ(a) when there exists a sequence (fn) in D(a) for which ‖fn‖ = 1and

limn→∞

(a− λ)fn = 0. (VI.3)

Of course, when λ is an eigenvalue of a with eigenvector f , we may take fn = f for all n.However, this is not the official definition of the spectrum, which is as follows.

Definition VI.3 Let a : D(a) → H be a (possibly unbounded) operator on a Hilbert space. Theresolvent ρ(a) is the set of all z ∈ C for which a− z : D(a) → H is injective and surjective (i.e.,invertible). The spectrum σ(a) of a is defined by σ(a) := C\ρ(a).

Hence the property (VI.1) has been turned into a definition! We will prove the equivalence ofthis definition of the spectrum with the definition above later on. In the example just given, one hasσ(a) = a(R) if the right domain of a is used, namely (V.1). Thus the spectrum can be nonemptyeven if there aren’t any eigenvalues. The subsequent theory shows that these are precisely the rightdefinitons for spectral theory.

The following result explains the role of closedness.2

Proposition VI.4 If an operator a : D(a) → R(a) = H has an inverse, then a−1 is bounded iff ais closed.

The proof consists of two steps. First, one has that a : D(a) → R(a) is closed iff a−1 is closed.To prove “⇒”, assume gn → g and a−1gn → f . Call fn := a−1gn; then afn = gn → g, so if ais closed then by definition f ∈ D(a) and afn → af , so af = g, hence f = a−1g, which meansa−1gn → a−1g. In particular, g ∈ R(a) = D(a−1), and it follows that a−1 is closed. The proof of“⇐” is the same, with a and a−1 interchanged. Geometrically, the graph of a−1 is just the imageof the graph of a in H ⊕ H under the map (f, g) 7→ (g, f), hence if one is closed then so is theother.

Secondly, if R(a) = H, then D(a) = H, hence a−1 is bounded by the closed graph theorem(Theorem V.5).

Returning to the equation (a − z)f = g, it now follows that when z ∈ ρ(a), the solutionf depends continuously on the initial data g iff a is closed. To avoid pathologies, we thereforeassume that a is closed in what follows. Furthermore, as we shall see, practically every argumentbelow breaks down when (a−z)−1 is unbounded. This also explains why as far as spectral theory isconcerned there isn’t much difference between bounded operators and closed unbounded operators:in both cases (a− z)−1 is bounded for z ∈ ρ(a).

As an exercise, one easily shows that for closed a one has

ρ(a∗) = ρ(a); σ(a∗) = σ(a). (VI.4)

Theorem VI.5 Let a be a closed operator.

1. ρ(a) is open (and hence σ(a) is closed) in C.

2. The operator-valued function z 7→ (a− z)−1 is analytic on ρ(a), i.e., it can be expressed as anorm-convergent power series in z.

2Some books define the resolvent of a as the set of those z ∈ C for which (a − z) is invertible and has boundedinverse. In that case, the resolvent is empty when a is not closed.

Page 39: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

VI.1. RESOLVENT AND SPECTRUM 39

Everything will follow from the following lemma, in which a ∈ B(H) is not the same a as inthe theorem!

Lemma VI.6 When ‖a‖ < 1 the sum∑nk=0 a

k converges in norm to (1− a)−1.

We first show that the sum is a Cauchy sequence. Indeed, for n > m one has

‖n∑k=0

ak −m∑k=0

ak‖ = ‖n∑

k=m+1

ak‖ ≤n∑

k=m+1

‖ak‖ ≤n∑

k=m+1

‖a‖k.

For n,m → ∞ this goes to 0 by the theory of the geometric series. Since B(H) is complete, theCauchy sequence

∑nk=0 a

k converges for n→∞. Now computen∑k=0

ak(1− a) =n∑k=0

(ak − ak+1) = 1− an+1.

Hence

‖1−n∑k=0

ak(1− a)‖ = ‖an+1‖ ≤ ‖a‖n+1,

which → 0 for n→∞, as ‖a‖ < 1 by assumption. Thus

limn→∞

n∑k=0

ak(1− a) = 1.

By a similar argument,

limn→∞

(1− a)n∑k=0

ak = 1.

so that, by continuity of multiplication of bounded operators (see (III.4)) one finally has

limn→∞

n∑k=0

ak = (1− a)−1.

Now let a be as in Theorem VI.5. Given z0 ∈ ρ(a), take the open set of all z ∈ C for which|z − z0|‖(a− z0)−1‖ < 1. Consequently,

(a−z0)−1∞∑k=0

(z − z0a− z0

)k= (a−z0)−1

(1− (z − z0)(a− z0)−1

)−1= (a−z0+(z0−z))−1 = (a−z)−1.

This proves existence of (a−z)−1, so that z ∈ ρ(a), which must therefore be open. Furthermore,it follows from the same calculation that (a− z)−1 is analytic in z.

Note that, as for ordinary (i.e., C-valued) analytic functions, analyticity implies continuity: with

‖(a− z)−1 − (a− z0)−1‖ = ‖∞∑k=1

(z − z0a− z0

)k‖ ≤ |z − z0|‖(a− z0)−1‖2

∞∑k=2

|z − z0|k‖(a− z0)−1‖k,

still assuming that |(z − z0|‖(a− z0)−1‖ < 1, one sees that limz→z0(a− z)−1 = (z − z0)−1.The following theorem is quite remarkable, since it shows that the norm of a bounded operator

on a Hilbert space is given purely algebraically, namely in terms of the spectrum.

Theorem VI.7 If a is bounded, then σ(a) is compact. Defining the spectral radius of a by

r(a) := sup|z|, z ∈ σ(a), (VI.5)

one has‖a‖ = r(a) (VI.6)

when a∗ = a, and‖a‖ =

√r(a∗a) (VI.7)

in general.

Page 40: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

40 CHAPTER VI. SPECTRAL THEORY FOR CLOSED UNBOUNDED OPERATORS

We will not prove this theorem here, as it is typical for bounded operators and our main interestin this course is in unbounded ones.3 It clearly implies that the spectrum of a bounded self-adjointoperator is never empty, and in fact we show in a different way that the spectrum of any boundedoperator on a Hilbert space is not empty:

Theorem VI.8 If a ∈ B(H), then σ(z) 6= ∅.

The idea of the proof is that when σ(z) = ∅ the operator-valued function z 7→ (a− z)−1, whichwe call g from now on, is not merely analytic on ρ(a) = C by Theorem VI.5, but in additionvanishes at infinity. Namely, for z 6= 0 we write ‖g(z)‖ = |z|−1‖(1 − a/z)−1‖ and observe thatlimz→∞ 1− a/z = 1, since limz→∞ ‖a/z‖ = 0. Hence limz→∞(1− a/z)−1 = 1, and

limz→∞

‖g(z)‖ = 0. (VI.8)

Since g is analytic and hence continuous, it follows that it is uniformly bounded (in z with respectto the operator norm). There is an operator-valued analogue of Liouville’s theorem from complexanalysis: any uniformly bounded analytic function on C taking values in the space of all boundedoperators on some Banach space is constant. By (VI.8) this constant is zero, so that g = 0. Thisis absurd, so that ρ(a) 6= C hence σ(a) 6= ∅.

On the other hand, for unbounded operators the spectrum can (literally) be any subset of C,including the empty set. The following example, continuing the example at the end of the previouschapter, describes some fairly unexpected possibilities.

Proposition VI.9 Let H = L2([0, 1]) and a = −id/dx : D(a) → H as a formal expression, withdomain to be specified below.

• If D(a) = f ∈ AC([0, 1]) | f ′ ∈ L2 & f(0) = 0 or D(a) = f ∈ AC([0, 1]) | f ′ ∈ L2 & f(1) =0, then σ(a) = ∅.

• If D(a) = f ∈ AC([0, 1]) | f ′ ∈ L2 & f(0) = f(1) = 0, then σ(a) = C.

• D(a) = f ∈ AC([0, 1]) | f ′ ∈ L2 & f(0) = f(1), then σ(a) = 2πZ.

Note that a is closed in all cases. In the first it is not symmetric, in the second it is symmetric,and in the third it is self-adjoint.

If z ∈ ρ(a), the differential equation

(−id/dx− z)g(x) = f(x) (VI.9)

should have a solution g ∈ D(a) for any f ∈ H, where the solution is obtained from f by a boundedoperator. This yields

g(x) = eizx(∫ x

0

dt e−iztf(t) +∫ 1

0

dt h(t)f(t)),

where h ∈ H is arbitrary (and independent of f) in so far as (VI.9) is solved; it should bechosen so as to satisfy the boundary condition inherent in D(a). Completing the argument is anexercise.

VI.2 The spectrum of self-adjoint operators

For a general closed operator a, we may decompose the spectrum as

σ(a) = σd(a) ∪ σc(a), (VI.10)

where the discrete spectrum σd(a) consists of all eigenvalues of a, and the continuous spec-trum σc(a) is the remainder of σ(a). Recall that eigenvalues lie in σ(a), for if (a−λ)f = 0 for somenonzero f then a−λ cannot be injective. The spectrum of self-adjoint operators has a particularlytransparent structure.

3See Weidmann Vol. I Ch. V or [15] for a proof.

Page 41: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

VI.2. THE SPECTRUM OF SELF-ADJOINT OPERATORS 41

Theorem VI.10 Let a be a self-adjoint operator (i.e., a∗ = a), and let z ∈ C. Then one of thefollowing possibilities occurs:

1. R(a− z) = H iff z ∈ ρ(a);

2. R(a− z)− = H but R(a− z) 6= H iff z ∈ σc(a);

3. R(a− z)− 6= H iff z ∈ σd(a).

The key to the proof is a very simple result.

Lemma VI.11 If a is closable (equivalently, if D(a∗) is dense), then R(a − z)− = N(a∗ − z)⊥

and N(a∗ − z) = R(a− z)⊥.

Note that the kernel of a closed operator (in this case a∗ − z) is automatically closed. Easycalculations using the definition of a∗ yield the inclusions R(a− z)⊥ ⊂ N(a∗ − z) and R(a− z) ⊂N(a∗ − z)⊥. Since K⊥⊥ = K− for any linear subspace K of a Hilbert space, and K ⊂ L impliesL⊥ ⊂ K⊥, the claim follows.

We first prove Theorem VI.10 for z ∈ R. If R(a− z)− 6= H, then N(a− z) = R(a− z)⊥ 6= 0, so(a− λ)f = 0 has a nonzero solution and λ ∈ σd(a). The converse implication has the same proof.If R(a − z)− = H, then N(a − z) = 0 and a − z is injective. Now if R(a − z) = H then a − zis surjective as well, and z ∈ ρ(a). The converse is trivial given the definition of the resolvent. IfR(a − z) 6= H, then z ∈ σc(a) by definition of the continuous spectrum. Conversely, if z ∈ σc(a)then z /∈ σd(a) and z /∈ ρ(a), so that R(a − z) = H and R(a − z)− 6= H are exlcuded by theprevious ‘iff’ results for ρ(a) and σd(a). Hence R(a− z)− = H but R(a− z) 6= H.

To prove Theorem VI.10 for z ∈ C\R, we first note that eigenvalues of self-adjoint operatorsmust be real; this is immediate since if a∗ = a then (f, af) = (af, f) = (f, af), so if f is aneigenvector with eigenvector λ it follows that λ = λ. In fact, we will prove that if z ∈ C\R, thenalso z ∈ ρc(a) is impossible, so that z ∈ ρ(a). To see this we need some lemma’s.

Lemma VI.12 Let a be symmetric. Then ‖(a− z)f‖ ≥ |Im (z)|‖f‖.

Reading Cauchy–Schwarz in the wrong order, we obtain

‖(a− z)f‖‖f‖ ≥ |(f, (a− z)f)| = |(r − iIm (z))‖f‖2| ≥ |Im (z)|‖f‖2.

Here we used the fact that r := (f, af) − Re (z) is a real number by virtue of the symmetry ofa.

Hence z ∈ C\R implies N(a − z) = 0. Combining this with Lemma VI.11, we infer thatR(a− z)− = N(a− z)⊥ = H. To infer that actually R(a− z) = H we need yet another lemma.

Lemma VI.13 Let a be any densely defined operator. If ‖af‖ ≥ C‖f‖ for some C > 0 and allf ∈ D(a), then a is injective and a−1 : R(a) → D(a) is bounded with bound ‖a−1‖ ≤ C−1.

Injectivity is trivial, for af = 0 cannot have any nonzero solutions given the bound; a linear mapa is injective when af = 0 implies f = 0. For the second claim, note that

‖a−1‖ = sup‖a−1g‖, g ∈ D(a−1) = R(a), ‖g‖ = 1 =

sup∥∥∥∥a−1 af

‖af‖

∥∥∥∥ , f ∈ D(a), f 6= 0

= sup∥∥∥∥ f

‖af‖

∥∥∥∥ , f ∈ D(a), f 6= 0.

This yields the claim.

Combining this with Lemma VI.12, we see that z ∈ C\R implies (a − z)−1 : D((a − z)−1) →D(a− z) = D(a) is bounded, where D((a− z)−1) = R(a− z). To infer that in fact R(a− z) = H,we use:

Lemma VI.14 If b is closed and injective, then b−1 : R(b) → D(b) is closed.

Page 42: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

42 CHAPTER VI. SPECTRAL THEORY FOR CLOSED UNBOUNDED OPERATORS

See the proof of Proposition VI.4.

Lemma VI.15 If b is closed and bounded, then D(b) is closed.

This is immediate from the definition of closedness.

Taking b = (a − z)−1, we find that D((a − z)−1) is closed. Since we know that R(a − z)− = H,we conclude that R(a − z) = H. The same is true for z. Hence by Lemma VI.11, N(a − z) =R(a− z)⊥ = H⊥ = 0 and a− z is injective. With a− z already known to be surjective, z ∈ ρ(a).

The proof of the converse implications is the same as for z ∈ R, and we have finished the proofof Theorem VI.10.

Using similar arguments, one can prove

Theorem VI.16 Let a be a symmetric operator. Then the following properties are equivalent:

1. a∗ = a, i.e., a is self-adjoint;

2. a is closed and N(a∗ ± i) = 0;

3. R(a± i) = H;

4. R(a− z) = H for all z ∈ C\R;

5. σ(a) ⊂ R.

Similarly, the following properties are equivalent:

1. a∗ = a∗∗, i.e., a is essentially self-adjoint;

2. N(a∗ ± i) = 0;

3. R(a± i)− = H;

4. R(a− z)− = H for all z ∈ C\R;

5. σ(a−) ⊂ R.

The second half of the theorem easily follows from the first, on which we will therefore concentrate.The implications 1 ⇒ 2, 1 ⇒ 4, 1 ⇒ 5 and 2 ⇒ 3 are immediate either from Theorem VI.10 orfrom its proof. The implications 4 ⇒ 3 and 5 ⇒ 4 are trivial. Thus it only remains to prove 3 ⇒ 1.

To do so, assume R(a ± i) = H. For given f ∈ D(a∗) there must then be a g ∈ H suchthat (a∗ − i)f = (a − i)g. Since a is symmetric, we have D(a) ⊂ D(a∗), so f − g ∈ D(a∗), and(a∗ − i)(f − g) = 0. But N(a∗ − i) = R(a + i)⊥ by Lemma VI.11, so N(a∗ − i) = 0. Hencef = g, and in particular f ∈ D(a) and hence D(a∗) ⊂ D(a). Since we already know the oppositeinclusion, we have D(a∗) = D(a). Given symmetry, this implies a∗ = a.

As an illustration of Theorem VI.16, one can directly show:

Proposition VI.17 Let a ∈ C(Ω) define a real-valued multiplication operator on

D(a) = f ∈ L2(Ω) | af ∈ L2(Ω) ⊂ H = L2(Ω),

so that a∗ = a (cf. Proposition V.9.) Then the operator a is injective iff a(x) 6= 0 for all x ∈ Ω,and surjective iff there exists ε > 0 so that |a(x)| ≥ ε for all x ∈ Ω; in that case a is injective andhas bounded inverse. Consequently, σ(a) = a(Ω)−, with a(Ω) := a(x), x ∈ Ω.

Finally, we justify our earlier heuristic definition of the spectrum; the thrust of the theoremlies in its characterization of the continuous spectrum, of course.

Theorem VI.18 Let a be self-adjoint. Then λ ∈ σ(a) iff there exists a sequence (fn) in D(a)with ‖fn‖ = 1 for all n such that limn(a− λ)fn = 0.

Page 43: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

VI.2. THE SPECTRUM OF SELF-ADJOINT OPERATORS 43

Suppose λ ∈ σ(a). If λ ∈ σd(a) we are ready, taking fn = f for all n. If λ ∈ σc(a), thenR(a−λ)− = H but R(a−λ) 6= H by Theorem VI.10. Now a is self-adjoint, hence a and a−λ areclosed, so that also (a−λ)−1 is closed by Lemma VI.14. Hence (a−λ)−1 : R(a−λ) → H must be a adensely defined unbounded operator by Lemma VI.15, for if it were bounded then its domain wouldbe closed, which D((a− λ)−1) = R(a− λ) is not, as we have just shown. Thus there is a sequencegn in D((a− λ)−1) with norm 1 and ‖(a− λ)−1gn‖ → ∞. Then fn := (a− λ)−1gn/‖(a− λ)−1gn‖has the desired property.

Conversely, if λ ∈ ρ(a) then (a− λ)−1 is bounded, hence (a− λ)fn → 0 implies fn → 0, so thesequence (fn) cannot exist, and λ ∈ σ(a) by reductio ad absurdum.

Page 44: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

44 CHAPTER VI. SPECTRAL THEORY FOR CLOSED UNBOUNDED OPERATORS

Page 45: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

Chapter VII

Quantum mechanics and Hilbertspace

We are now going to apply the previous machinery to quantum mechanics, referring to the Intro-duction for history and motivation. The mathematical formalism of quantum mechanics is easierto understand if it is compared with classical mechanics, of which it is a modification. We thereforestart with a rapid overview of the latter, emphasizing its mathematical structure.1

VII.1 Classical mechanics

The formalism of classical mechanics is based on the notions of phase space and time-evolution,going back to Descartes and Newton, and brought into its modern form by Hamilton. The phasespace of a given physical system is a collection of points, each of which is interpreted as a possiblestate of the system.2 At each instance of time, a given state is supposed to completely characterizethe ‘state of affairs’ of the system, in that:

1. The value of any observable (i.e., any question that may possibly be asked about the system,such as the value of its energy, or angular momentum,. . . ) is determined by it.3

2. Together with the equations of motion, the state at t = 0 is the only required ingredient forthe prediction of the future of the system.4

In general, the phase space M in classical mechanics is a so-called symplectic manifold. Ob-servables are then given by (preferably smooth) functions on M . In these notes we only look at thespecial case M = R2n, which describes a physical system consisting of a point particles moving inRn. We use coordinates (q, p) := (qi, pi), where i = 1, . . . , n. The q variable (“position”) denotesthe position of the particle, whereas the meaning of the p variable (“momentum”) depends onthe time-evolution of the system. For example, for a free particle of mass m one has the relation~p = m~v, where v is the velocity of the particle (see below). Let us note that one may look at, say,qi also as an observable: seen as a function on M , one simply has qi(q, p) = qi, etc.

Given the phase space M , the specification of the system is completed by specifying a (usuallysmooth) function h on M , called the Hamiltonian of the system. For M = R2n we therefore haveh as a function of (q, p), informally written as h = h(q, p). The Hamiltonian plays a dual role: it

1A more complete treatment should await a new course Symplectic Geometry and Classical Mechanics.2More precisely, as a pure state; see section VII.5 below.3Philosophers would say that any quantity pertaining to the system supervenes on its states; this means that no

change in a given quantity is possibly without a change in the state. For example, most scientists would agree thatthe mind supervenes on the brain (seen as a physical system).

4We do not say that such a prediction is always possible; see below. But if it is possible at all, it merely requiresthe state and the equations of motion.

45

Page 46: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

46 CHAPTER VII. QUANTUM MECHANICS AND HILBERT SPACE

gives the value of the energy, and determines the time-evolution of the system. Indeed, given hthe time-evolution is determined by Hamilton’s equations

qi :=dqi

dt=

∂h

∂pi;

pi :=dpidt

= − ∂h

∂qi. (VII.1)

For example, a particle with mass m moving in a potential V has Hamiltonian

h(q, p) =p2

2m+ V (q), (VII.2)

where p2 :=∑ni=1(pi)

2. The equations (VII.1) then read qi = pi/m and pi = −∂V/∂qi. Withthe force defined by F i := −∂V/∂qi, these are precisely Newton’s equations d2qi/dt2 = F i/m, or~F = m~a. In principle, h may explicitly depend on time as well.

The equations of motion (VII.1) may or may not have a solution for all times; this depends onboth h and the initial state (q, p). In the worst case, which occurs if the potential V in (VII.2)is sufficiently singular (e.g., not even Lipschitz5), then (VII.1) mat not even be solvable for smalltimes. The ssituation is slightly better when (VII.1) admits a solution for given (q, p) and |t| < ε,for some ε > 0. For example, if M = (0, 1)× R (think of a particle moving on a finite table) andh = p2/2m, then clearly the particle will roll off the table after some finite time; formally thismeans that (VII.1) has no solution for later times. In general, this situation applies when V isLipschitz (cf. [8]). But ideally, one expects the equations of motion to have a solution for all t ∈ R,given any initial state (q, p). In that case, the dynamics is said to be complete. Provided that hdoes not explicitly depend on t, it can be shown that a complete dynamics is given by an actionof R (seen as a group in the usual, additive way) on M .

Definition VII.1 An action of R on a set M is given by a map R× S → S, written as 〈t, σ〉 7→σ(t), satisfying σ(0) = σ and (σ(s))(t) = σ(s+ t) for all σ ∈ S and s, t ∈ R. The action is calledcontinuous or smooth when the map in question is continuous or smooth.

It is very hard to given conditions on M and h guaranteeing that the time-evolution is givenby a (continuous or smooth) action of R on M ; this will be much easier in quantum mechanics,as we will see below. In any case, it is only in that situation that we can say that the system isdeterministic. Thus the usual claim that classical mechanics is a deterministic theory (allegedly asopposed to quantum mechanics) is false!

VII.2 Quantum mechanics

Quantum mechanics is based on the postulate that the phase space is a Hilbert space H, with theadditional stipulations that:

1. Only vectors of norm 1 correspond to physical states;

2. Vectors differing by a “phase”, i.e., by a complex number of modulus 1, correspond to thesame physical state.

In other word, ψ ∈ H and zψ with z ∈ C and |z| = 1 give the same state.6 We here stick to thephysicists’ convention of denoting elements of Hilbert spaces by Greek letters.7

5A function f on Ω ⊂ Rn is called Lipschitz when there exists a constant C > 0 such that |f(x)−f(y)| ≤ C|x−y|for all x, y ∈ Ω. For example, any function with bounded first derivative is Lipschitz.

6It follows that the true state space of a quantum-mechanical system is the projective Hilbert space PH, whichmay be defined as the quotient SH/ ∼, where SH := f ∈ H | ‖f‖ = 1 and f ∼ g iff f = zg for some z ∈ C with|z| = 1.

7This notation was initially used by Schrodinger in order to make his wave mechanics, a precursor of quantummechanics, look even more mysterious than it already was.

Page 47: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

VII.3. THE SCHRODINGER EQUATION 47

The reason for the first point lies in the probability interpretation of quantum mechanics. Thesimplest example of this interpretation is given by the quantum mechanics of a particle moving inR3. In that case the Hilbert space may be taken to be H = L2(R3), and Born and Pauli claimed in1926 that the meaning of the ‘wavefunction’ ψ ∈ L2(R3) was as follows: the probability P (ψ, x ∈ ∆)that the particle in state ψ is found to be in a region ∆ ⊆ R3 is

P (ψ, x ∈ ∆) = (ψ, χ∆ψ) =∫

d3x ‖ψ(x)‖2. (VII.3)

Here χ∆ is the characteristic function of ∆, given by χ∆(x) = 1 when x ∈ ∆ and χ∆(x) = 0 whenx /∈ ∆. It follows that

‖ψ‖2 = (ψ,ψ) = P (ψ,Rn) = 1, (VII.4)

since by definition of the physical system in question we assume that the particle is somewhere.More generally, observables are represented in quantum mechanics by self-adjoint operators a

on H, and one interprets the real number (ψ, aψ) as the expectation value of the observable a inthe state ψ.8 In the next chapter, we will associate a projection p(∆) to each subset ∆ ⊂ σ(a) ofthe spectrum of a. The generalization of the Born–Pauli probability interpretation of ψ ∈ L2(R3)to arbitrary Hilbert space is then as follows: the probability P (ψ, a ∈ ∆) that the observable a takesvalues in ∆ ⊂ σ(a) ⊂ R when the system is in the state ψ ∈ H is given by

P (ψ, a ∈ ∆) = (ψ, p(∆)ψ). (VII.5)

As we shall see, p(σ(a)) = 1, so analogously to (VII.4) we may write

‖ψ‖2 = (ψ,ψ) = P (ψ, p(σ(a))ψ) = P (ψ, a ∈ σ(a)) = 1, (VII.6)

since in quantum mechanics one assumes a may only take values in its spectrum σ(a), and a hasto have some value. For example, a = 1 has σ(a) = 1 and p(1) = 1, so that P (ψ, 1 ∈ 1) =(ψ,ψ) = 1.

These considerations motivate rule 1 given at the beginning of this section; rule 2 follows fromthe fact that any conceivable prediction from quantum mechanics is given by a number of the form(ψ, aψ) for some ψ ∈ H and some self-adjoint operator a on H. Thus ψ and zψ (with |z| = 1)yields precisely the same predictions, and therefore correspond to the same state.

VII.3 The Schrodinger equation

Quantum mechanics has a single equation of motion, known as the Schrodinger equation. The roleof the Hamiltonian is now played by a certain operator h on a Hilbert space H, whose specificationis part of the definition of a given physical system. The precise form of this operator is usuallyguessed from the form of the classical Hamiltonian. For example, if H = L2(Rn) and the classicalHamiltonian is (VII.2), Schrodinger took the quantum Hamiltonian to be

h = − ~2

2m

n∑i=1

∂2

∂x2i

+ V (x). (VII.7)

Here ~ ∈ R+ is a constant of nature, called Planck’s constant, and V is seen as a multiplica-tion operator (as the notation already suggests). It is clear that h is an unbounded operator; weshall see shortly that it has to be self-adjoint for the whole theory to make sense. In general, if aclassical observable is a function f = f(qi, pi) on R2n, then roughly speaking the ‘corresponding’quantum observable is given by the operator f(xi,−i~∂/∂xi). This indeed yields (VII.7) from(VII.2), and even more simply says that the operator corresponding to (the i’th component of)

8This replaces the classical situation, where f(σ) is simply the sharp value of an observable f in a state σ. Whatthis really means is a matter of ongoing debate. For example, do expectation values refer to outcomes of a longseries of measurements? Or to ‘propensities’ of the observable to have the given expectation value? Or to averageswith respect to some unknown theory underlying quantum mechanics? Etcetera

Page 48: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

48 CHAPTER VII. QUANTUM MECHANICS AND HILBERT SPACE

position is xi (seen as a multiplication operator), and the operator corresponding to (the i’thcomponent of) momentum is −i~∂/∂xi. In general, however, the ‘quantization’ prescription

qi 7→ xi

pi 7→ −i~ ∂

∂xi(VII.8)

is only unambiguous when f does not contain products of q’s and p’s.In any case, given a quantum Hamiltonian h as an operator on some Hilbert space H, the

Schrodinger equation reads

hψ(t) = i~dψ(t)dt

. (VII.9)

This equation should be seen as the quantum-mechanical analogue of Hamilton’s equations (VII.1).As mentioned before, we shall establish below that h has to be self-adjoint on its domain D(h);clearly eq. (VII.9) only makes sense for ψ ∈ D(h). Nonetheless, the solutions of the Schrodingerequation should be defined for any ψ ∈ H, since any such ψ is allowed to be a physical state inquantum mechanics (as long as it has norm 1) and should have a well-defined time-evolution.

Let us write a solution of the Schrodinger equation with initial value ψ as

ψ(t) = u(t)ψ, (VII.10)

where u(t) : H → H is some map. We assume that we are in the ideal situation that the solutionis defined for all t ∈ R; we shall see immediately below that this is, in fact, always the case, inconstrast to the situation in classical mechanics. Indeed, if h is bounded (a situation that rarelyoccurs in practice except when H is finite-dimensional) and time-independent (which is true for aclosed system almost by definition), one may explicitly solve (VII.9) by (VII.10) with

u(t) = e−ith/~, (VII.11)

where the exponential is defined by its norm-convergent power-series. When h is not boundedthe exponential cannot be defined in this way, but provided it is self-adjoint the exponential canbe defined using the spectral theorem; see Chapter VIII. Eq. (VII.11) gives the solution of theSchrodinger equation also in that case.

VII.4 Stone’s theorem

Although we have not yet established (VII.11) in the physically relevant case where h is unbounded,some desirable properties of u(t) may be read off nonetheless. Taking these properties as axioms,we will then recover the notion of a Hamiltonian from Stone’s theorem below. First, as in classicalmechanics, when h is time-independent we expect that (ψ(s))(t) = ψ(s+t), and this indeed followsfrom (VII.11) and the formal computation exp(ta) exp(sa) = exp((s+t)a), where a is any boundedoperator.9 This translates into the property

u(s)u(t) = u(s+ t) (VII.12)

for all s, t ∈ R. Furthermore, one clearly has

u(0) = 0 (VII.13)

and the strong continuitylimt→0

u(t)ψ = ψ (VII.14)

for each ψ ∈ H; combined with (VII.12), this gives

limt→s

u(t)ψ = u(s)ψ (VII.15)

9The proof is the same as for the usual exponential function.

Page 49: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

VII.4. STONE’S THEOREM 49

for each s ∈ R. Finally, each u(t) given by (VII.11) is not only linear, as is immediate from (VII.11),and in general should be the case because the Schrodinger equation is linear, but also unitary.10

This follows from (VII.11) by a formal computation when h is bounded; in general, one could firstargue that u(t) has to be isometric because if ψ has norm 1, then also ψ(t) should have norm 1in order for the probability interpretation of quantum mechanics to make sense. Subsequently,(VII.12) and (VII.12) show that u(t) is invertible, with inverse u(t)−1 = u(−t). An easy exercisethen shows that an invertible isometric operator is unitary.

To some up, either intuition based on the Schrodinger equation or abstract arguments showthat time-evolution in quantum mechanics has the following structure. Recall Definition VII.1, inwhich the set M is now taken to be a Hilbert space H.

Definition VII.2 An action of R on a Hilbert space H of the form ψ(t) = u(t)ψ, where each u(t)is unitary and (VII.14) holds, is called a unitary representation of R on H.11

Note that the properties (VII.12) and (VII.13) are implicit in the definition of an action of R. Theunitary representation as a whole, given by the map t 7→ u(t), is usually simply called u.

On this basis, the relationship between the global time-evolution and the ‘infinitesimal’ time-evolution given by the Schrodinger equation is completely clarified by the following result.

Theorem VII.3 1. Let u be a unitary representation of R on H. Then the operator h, definedby the domain

D(h) : f ∈ H | limt→0

u(t)− 1t

f exists (VII.16)

and the action

hf := i limt→0

u(t)− 1t

f, (VII.17)

is self-adjoint. In particular, D(h) is dense in H.

In other words, provided ψ ∈ D(h), for each t ∈ R the vector ψ(t) = u(t)ψ (which lies inD(h) as well) satisfies the Schrodinger equation

hψ(t) = idψ(t)dt

. (VII.18)

2. Conversely, given a (densely defined) self-adjoint operator h on H there exists a uniqueunitary representation u of R on H that is related to h in the way just specified. Explicitly,one has

u(t) = e−ith, (VII.19)

where the exponential function is defined by the spectral calculus.

The last statement will be explained and proved in the next chapter. Of course, if one wants tointroduce Planck’s constant to recover (VII.9) and (VII.11), one should replace the operator h inthis theorem by h/~.

The first claim is proved as follows; we leave the details as an exercise. First, for any f ∈ Hand n ∈ N define

fn := n

∫ ∞

0

ds e−nsu(s)f.

Then fn ∈ D(h) and fn → f . Hence D(h) is dense in H. Second, an easy calculation shows thath is symmetric on D(h). Finally, one shows that R(h± i) = H and invokes Theorem VI.16.

10Recall that an operator u is unitary when uu∗ = u∗u = 1.11Condition (VII.14) or (VII.15) can be shown to be equivalent to the continuity of the map 〈t, ψ〉 7→ u(t)ψ from

R× H to H. When H is separable, one can show that in the presence of (VII.12) and (VII.13), condition (VII.14)is equivalent to to the requirement that the function t 7→ (f, u(t)g) is Lebesgue measurable from R to C for eachf, g ∈ H.

Page 50: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

50 CHAPTER VII. QUANTUM MECHANICS AND HILBERT SPACE

VII.5 States and observables

To close this chapter, we take a look at the formalism of quantum mechanics from a different per-spective. Any respectable physical theory known so far has the following mathematical structure:

• One has a state space S; this is a convex set (i.e., a subset of a given vector space, suchthat if ρ, σ ∈ S and λ ∈ (0, 1) then λρ+ (1− λ)σ ∈ S).

• One has a set of observables A; this is an ordered real vector space: i.e., there is aso-called positive cone A+ ⊂ A; this is a subset for which

1. a ∈ A+ and t ∈ R+ implies ta ∈ A+;

2. a, b ∈ A+ implies a+ b ∈ A+;

3. A+ ∩ −A+ = 0.

• One has a pairing S ×A→ R, written 〈ω, a〉 7→ ω(a), that is

1. convex in S, i.e., (λρ+ (1− λ)σ)(a) = λρ(a) + (1− λ)σ(a);

2. linear in A;

3. positive, in that ω(a) ≥ 0 for each a ∈ A+.

The interpretation of the pairing is that the observable a of the given system has value ω(a)in the state ω. In this sense, a state is just a value-assignment on the collection of all observablescharacterizing a physical system. However, it is not always clear what the ‘value’ of an observabledenotes. To clarify this point, we make a further assumption that is satisfied in both classicaland quantum mechanics, namely that the vector space of observables A has the structure of acommutative (but not necessarily associative algebra); in particular, that squares a2 of a ∈ A canbe defined. In that case, it turns out that

A+ = a2 | a ∈ A. (VII.20)

The variance of an observable a in a state ω is then defined by

∆ω(a) := ω(a2)− ω(a)2. (VII.21)

We say that an observable a has a sharp value in a state ω when ∆ω(a) = 0. In classical mechanicsthere exist states ω in which every observabel has a sharp value, but in quantum mechanicsthis is not the case due to the uncertainty principle (see below). If ∆ω(a) is nonzero, ω(a) isusually interpreted as the expectation value of a in ω. What this means in quantum mechanics iscontroversial; most practically minded people take it to be the average value of a with respect toa large number of measurements in which the system is prepared in the state ω each time.

Positivity of observables turns out to be an important ingredient in physical theories; think ofthe energy. It could be omitted from the axiomatic structure, but the notion of a state turns outto be clarified when it is included. An ordered real vector space may equivalently be characterizedby a linear partial ordering ≤, in which a ≤ b implies ta ≤ tb for each t ∈ R+ as well as a+c ≤ b+cfor each c ∈ A. Namely, given a positive cone one says that a ≤ b when b−a ∈ A+, and vice versa,given ≤ one defines A+ := a ∈ A | 0 ≤ a. For example, if A = C(Ω,R), i.e., the space al allcontinuous functions on Ω ⊆ Rn, then the subspace

A+ = C(Ω,R+) (VII.22)

of all positive functions is a positive cone in A. Of course, a positive function is the square ofits square-root, so that (VII.22) coincides with (VII.20). Also, if A = B(H)sa is the space of allbounded self-adjoint operators on some Hilbert space H, then

A+ := a2 | a∗ = a ∈ B(H) (VII.23)

Page 51: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

VII.5. STATES AND OBSERVABLES 51

is a positive cone, which turns out to coincide with the space of all operators on H with spectrumin R+, which in turn coincides with the space of all self-adjoint operators a for which (f, af) ≥ 0for all f ∈ H.

The convex structure of the state space reflects the idea of mixing or incomplete knowledge:given states ρ and σ, the state λρ + (1 − λ)σ, λ ∈ (0, 1) denotes the situation in which the stateof the system is unknown, but is estimated to be ρ with probability λ, and σ with probability1 − λ. Iterating this operation, one obtains states of the form ω =

∑i Piωi, where Pi > 0 and∑

i Pi = 1,12 in which one estimates that the system is in state ωi with probability Pi. Thus thestate space S can be decomposed into pure states and mixed states. A pure state is an elementω ∈ S for which ω = λρ+ (1− λ)σ for λ ∈ (0, 1) implies ρ = σ = ω. A mixed state is an elementof S that is not pure; hence mixed states do allow decompositions of the sort just mentioned.

Probably the simplest nontrivial example of a convex set is S = [0, 1], seen as a subset of R.The pure states in S are just 0 and 1. A more interesting example is the three-ball

B3 = (x, y, z) ∈ R3 | x2 + y2 + z2 ≤ 1; (VII.24)

as we shall see, this is the state space of a two-level system in quantum mechanics. Here the purestates are the points on the two-sphere

S2 = (x, y, z) ∈ R3 | x2 + y2 + z2 = 1. (VII.25)

However, its is not always the case that the set of pure states in a convex set S forms the geometricboundary of S. Indeed, a triangle ∆ with equal sides and equal angles is convex, but its pure statesare just its corners.13

The physical interpretation of a pure state is that it gives the most precise characterization ofthe given physical system that is possible in principle. Mixed states, on the other hand, have adirect interpretation only if their decomposition into pure states is unique; this depends on thegeometry of the state space. If so, mixed states have an ignornace interpretation as explainedabove: if ω =

∑i Piωi, where the ωi are pure, then the system is really in one of the states ωi. We

do not know which, but we assign probability Pi to the possibility ωi. In classical physics this isalways possible. On the other hand, this interpretation often fails in quantum physics, where thedecomposition of a mixed state into pure states is typically non-unique. For example, the originin B3 has an infinite number of possible decomopositions 0 = ω/2 + (−ω)/2, where ω ∈ S2 is anarbitrary pure state in B3. Indeed, it is doubtful that a quantum-mechanical system can be in apure state at all.14

General considerations go as far as what has been said above; for a specific physical theory onesimply has to state what the state space S, the set of observables A, as well as the pairing betweenthe two are. In classical mechanics one starts with a phase space M , as we have seen; just thinkof M = R2n, as above. One then puts

A = C∞b (S,R), (VII.26)

i.e., the set of all real-valued bounded smooth functions on M , with

A+ = C∞b (S,R+). (VII.27)

Furthermore,

S = P(M), (VII.28)

12Here the sum may even be infinite provided one has a suitable topology on S.13The sides of ∆ do have a geometric significance in convexity theory. A face F of a convex set S is a convex

subset of S that is closed under ‘purification’, in the sense that if ω ∈ F and ω = λρ+ (1− λ)σ for some λ ∈ (0, 1),then ρ ∈ F and σ ∈ F . Apart from ∆ itself and its pure states, the faces of ∆ are precisely its sides.

14This is due to the phenomena of entanglement and EPR-correlations.

Page 52: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

52 CHAPTER VII. QUANTUM MECHANICS AND HILBERT SPACE

the set of all probability measures on M . Here a probability measure on a topological space Mis an assignment B 7→ P (B) from suitable subsets B ⊂M to [0, 1],15 such that

P (∅) = 0; (VII.29)P (R) = 1, (VII.30)

P (∪iBi) =∑i

P (Bi) (VII.31)

for any countable collection of mutually disjoint sets Bi. Each point of R is interpreted as anevent, and the number P (B) associated to a subset B ⊂ R is interpreted as the probability thatsome event in the set B takes place. Such a measure leads to an integral

P (a) =∫dP (x) a(x) (VII.32)

for each a ∈ C∞(M), which defines the pairing between S and A. The pure states turn out to bethe points x of M , seen as probability measures Px by means of Px(B) = 1 if x ∈ B and Px(B) = 0if x /∈ B. Consequently, one has

Px(a) = a(x). (VII.33)

In quantum mechanics the model is quite different. One starts by choosing a Hilbert space Hthat bears some relationship to the underlying physical system.16 One then postulates

S = B1(H)+1 , (VII.34)

that is, the set of positive trace-class operators on H with unit trace (cf. Section IV.3). Elements ofB1(H)+1 are called density matrices by physicists, and are usually denoted by ρ. Since a densitymatrix is compact and self-adjoint, it is of the form ρ =

∑i λipi by Theorem IV.2. Recall that an

operator ρ ∈ B(H) is positive when (f, ρf) ≥ 0 for all f ∈ H. Positivity of ρ then yields λi ≥ 0 forall i, and Tr ρ = 1 is equivalent to

∑i λi = 1. In other words, a (bounded) operator ρ is a density

matrix iffρ =

∑i

Pipi, (VII.35)

where 0 < Pi ≤ 1 and∑i Pi = 1. It can be shown that the pure states are precisely the density

matrices with a single term, i.e., ρ = p for some one-dimensional projection p. In that case, ifpψ = ψ for a unit vector ψ, physicists would simply say that the system is in the pure state ψ(although even this seemingly straightforward statement is denied by some). In the general case,it is very natural to interpret the numbers Pi as probabilities, so that Pi is the probability thatthe system is in the pure state ωi. As already mentioned, this interpretation is actually quiteproblematic, as we shall now show in a very simple example.

Take H = C2. It can be shown (exercise) that a density matrix on C2 is of the form

ρ(x, y, z) :=12

(1 + x y + izy − iz 1− x

), (VII.36)

where (x, y, z) ∈ R3 satisfies x2 +y2 + z2 ≤ 1. This gives a bijection between the state space S andthe three-ball B3, which preserves the convex structure; in other words, as a convex set the statespace S of a two-level system can be seen as B3. As we have seen, the pure states form preciselythe geometric boundary S2 of B3. Referring back to out previous example, the point zero in B3

corresponds to the density matrix 1/2 (times the 2× 2 unit matrix). As we have seen, this densitymatrix admits uncountably many decompositions 1/2 · 1 = 1/2(ω1 + ω2) into pure states ωi: anypair of polar points on S2 will do.

15These are the so-called Borel subsets of M , which are countable unions or intersections of oen or closed subsetsof M .

16The specification of the Hilbert space plays a much weaker role in characterizing a quantum system than thespecification of a phase space in characterizing a classical system. In quantum mechanics the physical interpretationof the formalism comes from the procedure of quantization, which prescribes the meaning of each operator.

Page 53: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

VII.5. STATES AND OBSERVABLES 53

Let us return to general Hilbert spaces. The collection of observables in quantum mechanics is

A = B(H)sa, (VII.37)

the set of all bounded self-adjoint operators onH.17 The positive cone (VII.23) can be characterizedin several equivalent ways, for example, as

A+ = a ∈ B(H) | a∗ = a, σ(a) ⊂ R+. (VII.38)

Finally, the pairing between states and observables is given by

ρ(a) := Tr ρa. (VII.39)

Using (VII.35), we see thatρ(a) =

∑i

Pi(ψi, aψi), (VII.40)

where the ψi are unit vectors in piH. Thus ρ(a) ≥ 0 if a ≥ 0, as required.A key feature distinguishing quantum mechanics from classical mechanics is that even in pure

state the variance (VII.21) is typically nonzero; in fact, one has ∆ρ(a) = 0 iff ρ = p is pureand ψ (defined as above) is an eigenvector of a. The variance occurs in the famous Heisenberguncertainty relations. Noting that

∆ω(a) = ω((a− ω(a))2) ≥ 0,

we abbreviateσω(a) :=

√∆ω(a). (VII.41)

Theorem VII.4 • Let ρ ∈ B1(H)+1 be a density matrix on H, and let a, b ∈ B(H) be self-adjoint. Then (writing [a, b] = ab− ba for the commutator)

σρ(a)σρ(b) ≥12|ρ([a, b])|. (VII.42)

• Let a, b be symmetric operators, and let ψ ∈ D(a) ∩D(b) be a unit vector. Writing

σψ(a) =√

(aψ, aψ)− (ψ, aψ)2, (VII.43)

one hasσψ(a)σψ(b) ≥ |Im (aψ, bψ)|. (VII.44)

If ψ also lies in D(ab) ∩D(ba), this can be rearranged as as

σψ(a)σψ(b) ≥ 12|(ψ, [a, b]ψ)|. (VII.45)

The most famous application is the case a = qi = xi (as a multiplication operator) and b = pj =−i~∂/∂xj on H = L2(Rn); see(VII.8). Taking ψ ∈ C∞(Rn) ∩D(xi) for simplicity, with

D(xi) = ψ ∈ L2(Rn) |∫dnx (xi)2|ψ(x)|2 <∞,

we obtainσψ(qi)σψ(pj) ≥

~2δij . (VII.46)

17Most observables are actually represented by unbounded operators. Well! Nice axiomatics requires compromises.

Page 54: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

54 CHAPTER VII. QUANTUM MECHANICS AND HILBERT SPACE

Page 55: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

Chapter VIII

Spectral theory for self-adjointoperators

In this chapter we prove the spectral theorem for self-adjoint operators a. This theorem generalizesthe expansion a =

∑i λipi of a self-adjoint compact operator (such as a hermitian matrix), or

more generally of a bounded self-adjoint operator with purely discrete spectrum, in terms of itseigenvalues λi and the projections pi of the corresponding eigenspaces (i.e., if f ∈ piH thenaf = λif); see Theorem IV.7. As we have pointed out before, in general a self-adjoint operator neednot have any eigenvalues at all (recall the example of the multiplication operator a(x) = exp(−x2)on H = L2(R)), or may have mixed spectrum consisting of both eigenvalues and continuousspectrum (for example, a multiplication operator a ∈ C(Ω) on L2(Ω) with compact support haseigenvalue 0 with infinite multiplicity, since any f with support disjoint from a satisfies af = 0,whereas the remainder of the range of a will form its continuous spectrum).1

It turns out that an expansion of a in terms of “infinitesimal” projections p(t) = dE(t) (whichbecome genuine projections when t is an eigenvalue, or more generally when one integrates overt) and “generalized” eigenvalues t (that is, elements of σ(a)) is always possible. Such a “spectralexpansion” a =

∫dE(t) t:

• Gives a complete picture of a, including its domain;

• Enables one to define functions ϕ(a) of a by means of ϕ(a) =∫dE(t)ϕ(t), where ϕ : R → C

is an almost completely arbitrary function.2

In order to state the spectral theorem, we need some integration theory.

VIII.1 The Stieltjes integral

The notion of an integral, originally due to Newton, was made rigorous by Riemann in 1854. Hisdefinition was ∫ b

a

dx f(x) := limn∑i=1

∆xif(ξi),

where one has made a partition of [a, b] by n intervals Ii = [xi−1, xi] with distinguished pointsξi ∈ [xi−1, xi] and size ∆xi := xi − xi−1, and the limit is over a family of such partitions in whichthe size ∆xi goes to zero for all i. This definition is intuitively appealing, but it has two importantdrawbacks: it is limited to a small class of functions that can be integrated,3 and, even moreimportantly, it is tied to a specific way of assigning a “size” or “measure” ∆xi to a given interval

1For physicists: the spectrum of the Hamiltonian of the hydrogen atom has a mixed structure as well.2Well, is has to be Borel.3Consisting of the so-called Riemann-integrable functions, for which the limit exists. The class includes all

continuous and all monotonous functions, but not a great deal more.

55

Page 56: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

56 CHAPTER VIII. SPECTRAL THEORY FOR SELF-ADJOINT OPERATORS

[xi−1, xi]. For example, one would like to be able to use a “measure” which assigns a nonzero valueto a point; in the spectral theorem this will be necessary to deal with the discrete spectrum.

The modern notion of integration was introduced by Lebesgue in 1902, but we will not need thisgenerality; for the spectral theorem it turns out to be sufficient to use the definition of integrationdue to Stieltjes (1894). This definition is as appealing as the one of Riemann without sharing itsdrawbacks, and does not nearly have the level of abstraction of Lebesgue’s integral.4

The Stieltjes intergal (over R) is based on the notion of a density.

Definition VIII.1 A density is a function ρ : R → R that is:

1. monotone increasing, that is, ρ(s) ≤ ρ(t) if s ≤ t;

2. right-continuous, that is, limε→0+ ρ(t+ ε) = ρ(t).

In general, ρ need not and will not be left-continuous, but it follows from monotonicity that

ρ(t−) := limε→0+

ρ(t− ε) (VIII.1)

exists. For example, ρ(t) = c is a density, but it turns out that the resulting integral is identicallyzero. Much more interesting examples are

ρ(t) = t, (VIII.2)

which of course is both left- and right-continuous, and

ρ(t) = 0 (t < 1);ρ(t) = 1 (t ≥ 1). (VIII.3)

Now, given a density ρ, the associated Stieltjes integral∫dρ(t) f(t) of some function f : R → C

is first defined for simple functions f . By definition, these are of the form f =∑i ciχBi

, wherethe sum is finite, ci ∈ R, Bi ⊂ R is either an open interval Bi = (a, b) or a point Bi = a, allBi are mutually disjoint, and finally χB is the characteristic function of B (defined by χB(t) = 1when t ∈ B and χB = 0 when t /∈ B). For a simple function one puts∫

dρ∑i

ciχBi:=∑i

ci

∫dρ(t)χBi

, (VIII.4)

with ∫dρχ(a,b) := ρ(b−)− ρ(a);∫dρχa := ρ(a)− ρ(a−), (VIII.5)

from which it follows that ∫dρχ[a,b] = ρ(b)− ρ(a−); (VIII.6)∫dρχ[a,b) = ρ(b−)− ρ(a−); (VIII.7)∫dρχ(a,b] = ρ(b)− ρ(a). (VIII.8)

For an arbitrary positive function ϕ : R → R+ one then defines∫dρ(t)ϕ(t) ≡

∫dρϕ := sup

ψ simple, 0≤ψ≤ϕ

∫dρψ

. (VIII.9)

4Our presentation of Stieltjes integration will nevertheless borrow from Lebesgue’s later work.

Page 57: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

VIII.1. THE STIELTJES INTEGRAL 57

This supremum turns out to exist for a phenomenally large class of functions ϕ, namely the so-called Borel functions, and for this class the integral is defined by (VIII.9). If ϕ is bounded, it canbe shown5 that if a bounded monotone increasing sequence (ϕn) (i.e., one has

ϕ0 ≤ ϕ1 ≤ · · ·ϕn ≤ ϕn+1 ≤ · · · ≤ c

pointwise) converges pointwise to ϕ, then∫dρϕ = lim

n→∞

∫dρϕn. (VIII.10)

In particular, one may take each ϕn to be simple, for every bounded Borel function on R can beapproximated by such a sequence.

For arbitrary (Borel) functions ϕ : R → R one first decomposed ϕ = ϕ+ − ϕ−, where ϕ± ≥ 0,and defines ∫

dρϕ :=∫dρϕ+ −

∫dρϕ−. (VIII.11)

Finally, for ϕ : R → C one decomposes ϕ = ϕ′ + iϕ′′, where ϕ′ and ϕ′′ are real-valued, and defines∫dρϕ :=

∫dρϕ′ + i

∫dρϕ′′.

For example, for (VIII.2) one finds∫dρϕ =

∫R dt ϕ(t), and for (VIII.3) one obtains

∫dρϕ =

ϕ(1), something quite impossible with a Riemann integral! We will often use the calculation (whichis immediate from the definition):∫

dρχ(−∞,s] =∫dρχ(−∞,s) +

∫dρχs = ρ(s−)− ρ(−∞) + ρ(s)− ρ(s−) = ρ(s)− ρ(−∞).

(VIII.12)In the next section we encounter densities ρ with the properties

ρ(−∞) = 0;ρ(∞) = 1. (VIII.13)

We make this assumptions until the end of this section. In that case, ρ defines a probabilitymeasure on R (see Section VII.5) by

P (B) :=∫dρχB . (VIII.14)

For example, the density (VIII.3) satisfies (VIII.13), and with associated probabilities P (B) = 0if 1 /∈ B and P (B) = 1 if 1 ∈ B. In particular, P (1) = 1 and P (a) = 0 for all a 6= 1.

We may associate a Hilbert space L2(R, ρ) with such a probability measure, whose elements areequivalence classes of (Borel) functions ϕ : R → C for which

∫dρ(t)|ϕ(t)|2 is finite. It is possible

that ϕ 6= 0 but ‖ϕ‖ = 0. To exclude this possibility, the elements of L2(R, ρ) are taken to beequivalence classes [ϕ], where ϕ ∼ ψ when

∫dρ(t)|ϕ(t)−ψ(t)|2 = 0. The inner product in L2(R, ρ)

should strictly speaking be written as ([ϕ], [ψ]) =∫dρ(t)ϕ(t)ψ(t), where one should note that

the right-hand side is independent of the choice of the representatives f and g in the equivalenceclasses [ϕ] and [ψ], but we usually write this simply as

(ϕ,ψ) :=∫dρ(t)ϕ(t)ψ(t). (VIII.15)

It can be shown that the linear span of elements of the form χA, A ⊂ R as specified before, (orrather the equivalence classes these define) is dense in L2(R, ρ).

One should not be misled by the notation here, which falsely suggests a close analogy withL2(R) ≡ L2(R, id). For example, the function 1R is in L2(R, ρ), which is not the case for L2(R) ≡L2(R, id) (where id(t) = t; cf. (VIII.2)). More spectacularly, L2(R, ρ) need not even be infinite-dimensional! This completely depends on ρ. For example, with (VIII.3) one has (ϕ,ψ) = ϕ(1)ψ(1),

5Using the Lebesgue dominated convergence theorem

Page 58: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

58 CHAPTER VIII. SPECTRAL THEORY FOR SELF-ADJOINT OPERATORS

since∫dρ f = f(1) for any decent function f . Consequently, the map [ϕ] 7→ ϕ(1) from L2(R, ρ)

to C is isometric, and since it is clearly surjective it is unitary. Hence L2(R, ρ) ∼= C. In the nextsection we will encounter a density ρ for which

∫dρf =

∑ni=1 Pif(λi), where Pi > 0 with

∑i Pi = 1

and λi ∈ R. In that case, one has L2(R, ρ) ∼= Cn via the unitary map [ϕ] 7→√Piϕ(λi)ei, where

(ei) is the standard basis of Cn.

VIII.2 Spectral densities

For the spectral theorem a generalization of the Stieltjes integral is needed, in which the densitydoes not take vales in R but in the space of projections on a given Hilbert space H (namely theone on which the given self-adjoint operator a is defined).

Definition VIII.2 A spectral density on a Hilbert space H is a map E : R → B(H) for whicheach E(t) is a projection (i.e., E(t)∗ = E(t)2 = E(t)), and:

1. E is monotone increasing, that is, E(s) ≤ E(t) if s ≤ t in the sense that (f,E(s)f) ≤(f,E(t)f) for all f ∈ H or (equivalently) E(s)H ⊆ E(t)H;

2. E is strongly right-continuous, that is, limε→0+ E(t+ ε)f = E(t)f for each f ∈ H;

3. limt→−∞E(t)f = 0 and limt→∞E(t)f = f for each f ∈ H.

As for ordinary densities, it follows that although E may not be left-continuous, the limitlimε→0+ E(t− ε)f exists for all f .

As we shall see, each self-adjoint operator a on H defines a certain spectral density. To illustratethe latter concept, we run ahead of this and mention that for a multiplication operator a ∈ C(Ω)on H = L2(Ω), the associated spectral density is

E(t) = χa≤t := χx∈Ω|a(x)≤t, (VIII.16)

where χa≤t is, of course, seen as a multiplication operator onH (note that χB is always a projectionfor any decent B ⊂ Ω; the associated subspace χBH consists of all elements of L2 with support inB). As a special case, one may take a(x) = x on H = L2(R), in which case

E(t) = χ(−∞,t]. (VIII.17)

A completely different example is provided by the unit operator a = 1, which leads to

E(t) = 0 (t < 1);E(t) = 1 (t ≥ 1). (VIII.18)

Thirdly, if a =∑i∈I λipi, where pi projects on the eigenvector ei with eigenvalue λi and∑

i pi = 1, it turns out that

E(t) =∑

i∈I|λi≤t

pi. (VIII.19)

For an arbitrary spectral density E, each f ∈ H leads to a density ρf defined by

ρf (t) := (f,E(t)f), (VIII.20)

with an associated Stieltjes integral. In addition to the two properties mentioned in DefinitionVIII.1, in case that ‖f‖ = 1 one evidently has

3. limt→−∞ ρf (t) = 0 and limt→∞ ρf (t) = 1;

Page 59: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

VIII.2. SPECTRAL DENSITIES 59

cf. (VIII.13). In other words, if ‖f‖ = 1 then ρf defines a probability measure on R.For example, provided that ‖f‖ = 1, (VIII.18) leads to (VIII.3), with associated probabilities

as specified below (VII.30), and (VIII.17) leads to ρf (t) =∫ t−∞ dx |f(x)|2. This function happens

to be absolutely continuous in t (with derivative in L1), and the rules of Stieltjes integration implythat one may then compute integrals simply by

∫dρ(t) f(t) =

∫R dt dρ(t)/dtf(t). In the case at

hand one has dρf (t)/dt = |f(t)|2, so that the probabilities defined by ρf are

P (B) =∫dρχB =

∫Rdt χB(t)|f(t)|2 =

∫B

dt |f(t)|2. (VIII.21)

More generally, the spectral density E in (VIII.16) defined by a multiplication operator a onL2(Ω) leads to

ρf (t) =∫a≤t

dnx |f(x)|2, (VIII.22)

so that

P (B) =∫a−1(B)

dnx |f(x)|2, (VIII.23)

where a−1(B) = x ∈ Ω | a(x) ∈ B.Finally, the spectral density (VIII.19) yields

ρf (t) =∑

i∈I|λi≤t

Pi (VIII.24)

in terms ofPi := |(ei, f)|2. (VIII.25)

These numbers are precisely the probabilities associated to ρf , in the sense that

P (B) =∑

i∈I|λi∈B

Pi, (VIII.26)

so that in particular one hasP (λi) = |(ei, f)|2 (VIII.27)

and P (a) = 0 if a 6= λi for some i ∈ I.The probabilities thus defined by a given spectral density on H and a unit vector f ∈ H may

be written asP (B) = (f, p(B)f), (VIII.28)

where

p(B) :=∫dE χB (VIII.29)

for a subset B ⊂ R that is a finite union of points and open intervals; the definition of∫dE χB is

exactly the same as (VIII.4) and (VIII.5), with ρ replaced by E. It is easily seen that each p(B)is a projection, and that

1. p(∅) = 0 and p(R) = 1;

2. p(A ∩B) = p(A)p(B);

3. p(A ∪B) = p(A) + p(B) when A ∩B = ∅;

4. p(∪nBn) = ∨p(Bn), where ∨pn is the smallest projection p such that pn ≤ p for all n.

Page 60: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

60 CHAPTER VIII. SPECTRAL THEORY FOR SELF-ADJOINT OPERATORS

It follows from the second item that all p(B) commute among each other when B varies.p(Bi)p(Bj) = 0 if Bi ∩Bj = ∅. The association B 7→ p(B) is called the spectral measure on

R defined by the spectral density E; it is a projection-valed probability measure, which is of greatimportance for quantum mechanics.

In case that E is defined by a self-adjoint operator a, the points on R are interpreted as possiblevalues a can take upon measurement, and one has the following generalization of the Born–Pauliprobability interpretation of quantum mechanics:

If ψ is a unit vector in H, representing the state of the system, and a is a self-adjoint operator,representing some physical observable, then the probability that a takes some value in B ⊂ R is

Pψ(a ∈ B) = (ψ, p(B)ψ), (VIII.30)

where p(B) is defined by (VIII.29) in terms of the spectral density E associated to a.

Physicists will recognize that the special cases so far agree with what they know from quantummechanics courses; (VIII.21) is just the Pauli interpretation of the wave function, and (VIII.27) isthe Born interpretation. As we shall see, ρf turns out to be supported by the spectrum σ(a); thismeans that P (B) = 0 whenever B ∩σ(a) = ∅. Thus a self-adjoint operator can only assume valuesin its spectrum.

To be honest, we have only recovered the Pauli interpretation in one dimension. To do sofor a particle moving in Rn, we need to extend the probability interpretation just stated to thecase of more than one spectral density. This extension was first given by von Neumann [19], asfollows. Suppose one has n commuting self-adjoint operators ak, with associated spectral densitiesEk. It can be shown (cf. [22]) that all Ek(t) commute among each other for all values of t; thisis a necessary condition for the consistency of the probability interpretation. The analogue of thecase of a single operator is:

The probability that in some state ψ the observables (a1, . . . , an) simultaneously take some valuein B1 × · · · ×Bn is

Pψ(a1 ∈ B1, . . . , an ∈ Bn) = (ψ, p(B1) · · · p(Bn)ψ), (VIII.31)

where p(Bi) is defined by (VIII.29) with E = Ei, the spectral density associated to ai.

For example, if H = L2(Rn) and ak = xk as a multiplication operator, then Ek(t) = χ(−∞,tk]

and hence p(Bk) = χBk(xk) as a multiplication operator. Hence

Pψ(x1 ∈ B1, . . . , xn ∈ Bn) =

∫B1×···×Bn

dnx |ψ(x)|2. (VIII.32)

Eq. (VIII.17) is more than just a special case of a spectral density: it is to some extent thegeneral case. To see this, given some spectral density E on a Hilbert space H we pick a vectorf ∈ H with ‖f‖ = 1, and define

Hf := spanE(t)f, t ∈ R−. (VIII.33)

This is a closed subspace of H. We then define uf : spanE(t)f, t ∈ R → L2(R, ρf ) by linearextension of

uf : E(t)f 7→ χ(−∞,t]. (VIII.34)

In particular, one hasufp(B)f = χB ; (VIII.35)

for example, uf (f) = 1R. An easy calculation (exercise!) shows that uf is isometric, so that (beingbounded, in particular) it can be extended to an operator uf : Hf → L2(R, ρf ). Furthermore, werecall that the set of all step functions is dense in L2(R, ρf ), so that uf (spanE(t)f, t ∈ R) is densein L2(R, ρf ). It follows (exercise) that uf is unitary. Now, it trivially follows from (VIII.34) thatuE(s)E(t)f = χ(−∞,s]uE(t)f ; if s < t the left-hand side is uE(s)f = χ(−∞,s], and the right-hand

Page 61: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

VIII.3. OPERATOR-VALUED STIELTJES INTEGRALS 61

side is χ(−∞,s]χ(−∞,t] = χ(−∞,s] as well. If s ≥ t one similarly finds χ(−∞,t] = χ(−∞,t]. Since uf isunitary, and the span of all E(t)f is dense in Hf by definition, one finds ufE(s)u−1

f = χ(−∞,s] asa multiplication operator on L2(R, ρf ) (whereas the right-hand side of (VIII.34) is an element ofL2(R, ρf )). In particular, one ufp(B)u−1

f = χB as a multiplication operator on L2(R, ρf ) for anyB ⊂ R (of the stated sort).

If Hf = H, we have now succeeded in showing that any spectral density is unitarily equivalentto (VIII.17), though defined on L2(R, ρf ) rather than on L2(R). If Hf 6= H, we write f = f1 andpick f2 ∈ H⊥

f and repeat the procedure. With H separable, we eventually arrive at H = ⊕nHfn ,and obtain a unitary operator

u = ⊕nufn : H →⊕n

L2(R, ρfn) (VIII.36)

by linear extension of ufn: E(t)fn 7→ χ(−∞,t], where the right-hand side is seen as an element of

L2(R, ρfn). As above, one obtains

uE(s)u−1 = χ(−∞,s], (VIII.37)

where the right-hand side is a single multiplication operator on ⊕nL2(R, ρfn) (i.e., on any g ∈L2(R, ρfn) one has uE(s)u−1g = χ(−∞,s]g), and

up(B)u−1 = χB . (VIII.38)

Consequently, any spectral density is unitarily equivalent to (VIII.17), defined as a multiplicationoperator on ⊕nL2(R, ρfn

).

VIII.3 Operator-valued Stieltjes integrals

Given a spectral density E on a Hilbert space H and a (Borel) function ϕ : R → C, we are nowgoing to define an operator

ϕ(E) =∫dE(t)ϕ(t) (VIII.39)

on the domain

Dϕ(E) := f ∈ H | ϕ ∈ L2(R, ρf ) = f ∈ H |∫dρf (t) |ϕ(t)|2 <∞. (VIII.40)

The definition will be such that if ϕ =∑i ciχBi

is simple, then∑i

ciχBi(E) ≡∫dE

∑i

ciχBi =∑i

cip(Bi), (VIII.41)

where p(B) was defined in (VIII.29). Note, in any case, that it is a highly nontrivial fact (exercise)that Dϕ(E) is a dense linear subspace of H.

For any bounded function ϕ one has ϕ ∈ L2(R, ρf ) for all f ∈ H by (VII.30), so Dϕ(E) = H,hence one expects ϕ(E) to be a bounded operator. Indeed, when ϕ is simple (and hence bounded)the same calculation that establishes the unitarity of the map uf in the previous section yields

‖ϕ(E)f‖ = ‖ϕ‖L2(R,ρf ). (VIII.42)

Since‖ϕ‖2L2(R,ρf ) =

∫dρf |ϕ(t)|2 ≤ ‖ϕ‖2∞(ρf (∞)− ρf (−∞)) = ‖ϕ‖2∞,

we see that ϕ(E) is bounded for simple ϕ. We now first define ϕ(E) for a bounded function ϕ,and will find the same conclusion.

Page 62: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

62 CHAPTER VIII. SPECTRAL THEORY FOR SELF-ADJOINT OPERATORS

Suppose that ϕ ≥ 0 and let (ϕn) be a bounded monotone increasing sequence (ϕn) of simplefunctions converging pointwise to ϕ; cf. (VIII.10) and surrounding text. Such a sequence exists bya deep fact of real analysis. Inspired by (VIII.10), we put

ϕ(E)f := limnϕn(E)f (VIII.43)

for each f ∈ H, where ϕn(E) is defined by linearity and (VIII.29), that is,∑k

ckχBk(E) =

∑k

ckp(Bk), (VIII.44)

where p(B) is given in (VIII.29).To see that this definition makes sense, we introduce and important proposition, whose proof

we leave as an exercise. Notationwise, if a and b are bounded self-adjoint operators, a ≤ b bydefinition means that (f, af) ≤ (f, bf) for all f ∈ H

Proposition VIII.3 Each bounded sequence (an) of monotone increasing bounded self-adjointoperators (i.e., one has

a0 ≤ a1 ≤ · · · an ≤ an+1 ≤ · · · ≤ c1 (VIII.45)

for some constant c; equivalently, ‖an‖ ≤ c for all n) converges strongly (that is, in the sense that(anf) converges in H for each f ∈ H) to a bounded self-adjoint operator a, which is the least upperbound of the given sequence (i.e., an ≤ a for all n, and if an ≤ b for all n then a ≤ b).

This proposition should be compared with the analogous result for real numbers, which isindeed a key step in its proof. Now, it is easy to show (exercise), yet of great importance, that:

Lemma VIII.4 If (ϕn) is a monotone increasing sequence (ϕn) of simple functions bounded byc, then the sequence of operators (an := ϕn(E)) satisfies (VIII.45) for the same value of c.

Consequently, (VIII.43) makes good sense by virtue of Proposition VIII.3. In particular, ϕ(E)is bounded for bounded ϕ, as expected. For general ϕ, we then define the integral ϕ(E) bydecomposing ϕ into two (in the real case) of four (in the complex case) positive parts, as in thedefinition of the ordinary Stieltjes integral.

To define ϕ(E) for arbitrary, i.e., not necessarily bounded (Borel) functions, we need a slightymore subtle procedure. We once again use the fact that the collection of simple functions is dense inL2(R, ρf ) for each E and each f ∈ H. If now ϕ ∈ L2(R, ρf ), so that f ∈ Dϕ(E) (see (VIII.40)), wepick a sequence of simple functions ϕn converging to ϕ in L2(R, ρf ), and define ϕ(E) by (VIII.43)once more. To see that this limit exists, we note that (ϕn) is a Cauchy sequence in L2(R, ρf ), sothat (ϕn(E)f) is a Cauchy sequence in H, since

‖ϕn(E)f − ϕm(E)f‖H = ‖ϕn − ϕm‖L2(R,ρf ).

In this limiting procedure one may well lose the boundedness of ϕ(E) for simple ϕ.6 Nonetheless,from (VIII.42) (established for simple functions ϕ) and the definition of ϕ(E) one concludes that(VIII.42) holds for arbitrary (Borel) functions ϕ, as long as f ∈ Dϕ(E), so that the right-handside (and consequently the left-hand side) is finite. Moreover, one has a very simple criterion forself-adjointness:

Proposition VIII.5 The operator ϕ(E) defined by (VIII.39) and (VIII.40) has adjoint

ϕ(E)∗ = ϕ(E), (VIII.46)

where ϕ is the complex conjugate of ϕ. In particular, ϕ(E) is self-adjoint when ϕ is real-valued.

6If ϕ is bounded, the two definitions we have given of ϕ(E) are consistent, since by the Lebesgue dominatedconvergence theorem one has ϕn → ϕ in L2(R, ρf ) for all f ∈ H. Hence the second definition includes the first, andin fact enables one to compute ϕ(E) using a larger class of approximants that the first definition.

Page 63: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

VIII.4. THE SPECTRAL THEOREM FOR BOUNDED OPERATORS 63

We leave the proof as an exercise.An equivalent way to define ϕ(E) for general ϕ is to regard ϕ as a (possibly unbounded)

multiplication operator on⊕

n L2(R, ρfn), and then put

ϕ(E) = u−1ϕu, (VIII.47)

where u was defined around (VIII.36); cf. (VIII.38). To be precise, let ϕn be ϕ as a multiplicationoperator on Hn = L2(R, ρfn

) (where fn ∈ Dϕ(E)), which is defined and self-adjoint on

D(ϕn) = ψ ∈ L2(R, ρfn) | ϕψ ∈ L2(R, ρfn

). (VIII.48)

We then use the following result (exercise):

Proposition VIII.6 Let H =⊕

nHn and suppose that for each n one has a self-adjoint operatoran on Hn. Then there is a unique self-adjoint operator a on H whose restriction to Hn is an,namely the operator defined on the domain

D(a) = f =∑n

fn | fn ∈ D(an) &∑n

‖anfn‖2 <∞ (VIII.49)

by a∑n fn :=

∑n anfn.

This makes it clear what the domain of the multiplication operator ϕ on⊕

n L2(R, ρfn

), andhence provides a complete interpretation of (VIII.47).

VIII.4 The spectral theorem for bounded operators

The central result of this chapter is Theorem VIII.12 in the next section, but for pedagogicalreasons we first state and prove the bounded case (to which the proof of the general case will evenbe reduced!7).

Theorem VIII.7 Let a be a bounded self-adjoint operator on H. Then there exists a uniquespectral density E with the property

a = id(E) =∫dE(t) t. (VIII.50)

We first construct the E(t) from a. To do so, we are going to define a self-adjoint operatorϕ(a) for each bounded real-valued (Borel) function ϕ on R. This procedure will be analogous tothe definition of ϕ(E): the definition will first be given for a class of functions ϕn for which itis a priori clear what ϕn(a) should mean, and the general case will be handled by approximatingan arbitrary ϕ by a pointwise convergent monotone sequence of the ϕn in the given clas. Thedifference between the construction of ϕ(E) and that of ϕ(a) is that in the former case the ϕn aresimple functions, whereas in the latter case they are polynomials.

So let a be a bounded self-adjoint operator on H. For a polynomial p(t) =∑k ckt

k it is clearthat p(a) should equal

p(a) =∑k

ckak. (VIII.51)

This is again a bounded operator by (III.4), which is self-adjoint if ck ∈ R. In particular,

id(a) = a, (VIII.52)

where id(t) = t. Note that p(a) is a bounded operator despite the fact that p is generally unboundedas a function on R. The reason for this is that a satisfies a bound α11 ≤ a ≤ a21, where αi ∈ R;it follows from the Cauchy–Schwarz inequality that this is certainly the case for α1 = −‖a‖ and

7For a uniform but somehwat elaborate proof for the bounded and unbounded case at one stroke see [26]. Ourproof combines ideas from [24] and [20].

Page 64: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

64 CHAPTER VIII. SPECTRAL THEORY FOR SELF-ADJOINT OPERATORS

α2 = ‖a‖, but we may take α1 = inf(f, af), ‖f‖ = 1 and α2 = sup(f, af), ‖f‖ = 1 (one ofthese will equal ±‖a‖). We write Ia := [α1, α2]. It turns out that in defining ϕ(a) only the values ofϕ on Ia are relevant, so that we may regard ϕ as a function from Ia to R. Clearly, any polynomial(or, more generally, any continuous function) is bounded on Ia.

It is of central importance for what follows that the analogue of Lemma VIII.4 holds:

Lemma VIII.8 If (ϕn) is a monotone increasing sequence (ϕn) of real polynomials on Ia boundedby c, then the sequence of operators (an := ϕn(a)) satisfies (VIII.45) for the same value of c.

Once again, we leave the proof of this lemma as an exercise. It shows, for example, that one cannotrestrict the polynomials p to an interval smaller than Ia, for if p(t) ≤ q(t) merely on such a smallerinterval, then one may not necessarily have p(a) ≤ q(a).

Similarly, one has

Lemma VIII.9 If pq is the pointwise product of two polynomials p and q, then (pq)(a) = p(a)q(a),and similarly for p + q and cp, c ∈ R. Moreover, one has p(a)∗ = p(a) (where p is the complexconjugate of p). Finally, one has

‖p(a)‖ ≤ ‖p‖∞ = supt∈Ia

|p(t)|. (VIII.53)

Now let ϕ : Ia → R+ be a bounded positive (Borel) function. It is a deep fact of analysis (see,e.g., [20]) that there exists a bounded monotone increasing sequence (ϕn) of polynomials on Iathat converges pointwise to ϕ. Consequently, Proposition VIII.3 enables us to define

ϕ(a)f := limnϕn(a)f (VIII.54)

for each f ∈ H, where ϕn(a) is defined by (VIII.51). For arbitrary (Borel) ϕ : Ia → C, one thendefines ϕ(a) by decomposing ϕ as a sum of four positive parts.

Thus, for a given bounded self-adjoint operator a, we have defined an operator ϕ(a) for eachbounded Borel function ϕ : Ia → C, which is is self-adjoint if ϕ is real-valued. For fixed a thecorrespondence a 7→ ϕ(a) may also be reagrded as a map ϕ 7→ ϕ(a) from the space Bb(Ia) ofbounded Borel functions on Ia to B(H) (where H is the Hilbert space on which a is defined). Thismap preserves all good properties of its restriction to polynomials. To state these, note that Bb(Ia)is an algebra under pointwise operations. In particular, one has

Proposition VIII.10 Let a be a bounded self-adjoint operator on H, with Ia as above.

1. The map ϕ 7→ ϕ(a) from the space Bb(Ia) of bounded (Borel) functions on Ia to B(H) is analgebra homomorphism, which in addition satisfies ϕ(a)∗ = ϕ(a).

2. If (ϕn) is a bounded monotone sequence of (Borel) functions on Ia converging pointwise toϕ, then ϕ(a) = limn ϕn(a) (strong limit, i.e., ϕ(a)f = limn ϕn(a)f for each f ∈ H).

3. One has‖ϕ(a)‖ ≤ ‖ϕ‖∞. (VIII.55)

This proposition is sort of obvious given Lemmas VIII.8 and VIII.9 along with the definition ofϕ(a) by (VIII.54); we refer to [20] (Thm. 4.5.4) for a proof.

After this preparation, we define E(t) in terms of a by

E(t) = 0 (t < α1);E(t) = χ[α1,t](a) (t ∈ [α1, α2]);E(t) = 1 (t > α2). (VIII.56)

Using Proposition VIII.10, it is easily shown that E satisfies the axioms of a spectral density. Forexample, each E(t) is a projection because (for t ∈ [α1, α2])

E(t)2 = χ[α1,t](a)2 = χ2

[α1,t](a) = χ[α1,t](a) = E(t).

Page 65: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

VIII.4. THE SPECTRAL THEOREM FOR BOUNDED OPERATORS 65

Furthermore, since χ[α1,t] is right-continuous as a function of x, the right-continuity of E(t) followsfrom the second part of the proposition.

Recall (VIII.29), and notice thatp(B) = χB(a) (VIII.57)

for B ⊂ R (restricted as before) by (VIII.44) and (VIII.56). Here χB(a) := χB∩Ia(a), since χB(a)is not defined as such.8 For example, for x, y ∈ Ia one has

χ(x,y](E) = E(y)− E(x) = χ[α1,y](a)− χ[α1,x](a) = (χ[α1,y] − χ[α1,x])(a) = χ(x,y](a).

The preceding calculation shows that

ϕn(E) = ϕn(a) (VIII.58)

for all simple functions ϕn. Hence if ϕn → ϕ in the stated way (i.e., pointwise and monotoneincreasing), then (VIII.58), (VIII.43), and Proposition VIII.10.2 yield

ϕ(E) = limnϕn(E) = lim

nϕn(a) = ϕ(a),

where both limits are strong. In other words (cf. (VIII.39)),

ϕ(E) = ϕ(a). (VIII.59)

Clearly, (VIII.50) follows as a special case.The uniqueness of E may be derived from Stone’s formula:

ρf (t) = limδ→0+

limε→0+

∫ t+δ

−∞

dx

2πi(f, [(a− x− iε)−1 − (a− x+ iε)−1]f), (VIII.60)

which determines ρf and hence E (cf. (VIII.20))9 in terms of a, given (VIII.59). Stone’s formulafollows from Stieltjes’ formula

ρ(t) = limδ→0+

limε→0+

∫ t+δ

−∞

dx

2πi[g(x+ iε)− g(x− iε)], (VIII.61)

where, for z ∈ C\R,

g(z) :=∫ ∞

−∞dρ(t)

1t− z

. (VIII.62)

Namely, putting ϕ : t 7→ (t− z)−1 in VIII.59 one finds

1a− z

=∫dE(t)

1t− z

(VIII.63)

and hence, for ρ = ρf ,g(z) = (f, (a− z)−1f). (VIII.64)

Eq. (VIII.61) then immediately yields (VIII.60).10

This finishes the proof of Theorem VIII.7.

The spectral theorem is of fundamental importance to quantum mechanics, where the spectraldensity E defined by a, rather than a itself determines the probability interpretation of the the-ory; see (VIII.30) and (VIII.31). More generally, E is important because it contains all spectralinformation about a. We state the result also for unbounded operators, as we will see in the nextsection that the proof below (for the bounded case) also implies the result in general.

8It is possible to define E by (VIII.57), because p determines E just as well as E determines p.9If b is any bounded self-adjoint operator, then (f, bf) = 0 for all f ∈ H implies b = 0. Hence b is determined by

the collection of all (f, bf).10Stone’s formula may be used to define E in terms of a, see [26], but although this leads to a unified proof of the

spectral theorem for bounded and unbounded operators, it gives little insight in the matter.

Page 66: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

66 CHAPTER VIII. SPECTRAL THEORY FOR SELF-ADJOINT OPERATORS

Theorem VIII.11 Let a be a self-adjoint operator, with associated spectral density E (see (VIII.56))and projection-valued measure p (cf. (VIII.57) and (VIII.29)).

1. One has λ ∈ σ(a) iff p(λ− ε, λ+ ε) 6= 0 for all ε > 0.

2. One has λ ∈ σd(a) iff p(λ) 6= 0, in which case p(λ) equals the projection onto theeigenspace Hλ of a.11

The implication “⇒” in part 1 is logically equivalent to: if p(λ− ε, λ+ ε) = 0 for some ε > 0,then λ ∈ ρ(a), i.e., (a− λ)−1 exists (as a bounded operator). Indeed, define

αλ(t) =1

t− λ; t /∈ (λ− ε, λ+ ε);

αλ(t) = 0; t ∈ (λ− ε, λ+ ε)

Then ‖αλ‖∞ ≤ ε andαλ(id− λ) = (id− λ)αλ = 1 (VIII.65)

on R\(λ− ε, λ+ ε). Now, if p(λ− ε, λ+ ε) = 0, then the definition of the operator-valued Stieltjesintegral implies that ∫

dE ψ =∫dE χR\(λ−ε,λ+ε)ψ (VIII.66)

for all reasonable (e.g., bounded Borel) functions ψ. By (VIII.59) for ϕ = αλ(id−λ) (and similarlyfor ϕ = (id− λ)αλ) one has

(αλ(id− λ))(a) =∫dE(t)αλ(t)(t− λ) = 1 (VIII.67)

by (VIII.66) and (VIII.65). By Proposition VIII.10.1 the left-hand side equals αλ(a)(a−λ), henceαλ(a)(a− λ) = 1, and similarly (a− λ)αλ(a) = 1. Hence αλ(a) = (a− λ)−1, so that λ ∈ ρ(a).

In the “⇐” direction, take ε = 1/n and pick some fn ∈ p(λ − 1/n, λ + 1/n)H with norm 1;such f exists by assumption. Then, using by Proposition VIII.10.3 and (VIII.57),

‖(a−λ)fn‖ = ‖(a−λ)p(λ−1/n, λ+1/n)fn‖ ≤ ‖(a−λ)p(λ−1/n, λ+1/n)‖ ≤ ‖(id−λ)χ(λ−1/n,λ+1/n)‖∞.

The right-hand side is ≤ 1/n, hence ‖(a− λ)fn‖ → 0. By Theorem VI.18, λ ∈ σ(a).To prove part 2, we write p(λ) ≡ p(λ) and denote the projection onto the eigenspace Hλ by

pλ (provided λ ∈ σd(a)). We have p(λ) = χλ(a) by (VIII.29); we write χλ ≡ χλ. Hence

ap(λ) = id(a)χλ(a) = (idχλ)(a) = (λ1χλ)(a) = λ1(a)χλ)(a) = λp(λ).

Consequently, if p(λ) 6= 0 then pλ 6= 0, so that λ ∈ σd(a), and

p(λ) ≤ pλ, (VIII.68)

since p(λ)H ⊆ pλH = Hλ, as we have just shown.Conversely, let ϕn(t) := (inf1, |λ − t|)1/n, so that ϕn → χR\λ pointwise. Hence by Propo-

sition VIII.10.2 one has ϕn(a) → χR\λ(a) = p(R\λ). Now, if λ ∈ σd(a), so that pλ 6= 0,then

ϕn(a)pλf = ϕn(λ)pλf = 0 (VIII.69)

for all f ∈ H; for in general one has ϕ(a)pλf = ϕ(λ)pλ, as can be seen by approximating ϕby polynomials (for which the property is clearly true) and using Proposition VIII.10.2. Lettingn → ∞ in (VIII.69) yields p(R\λ)pλ = 0, or p(λ)pλ = pλ, as p(R\λ) + p(λ) = 1. In otherwords,

p(λ) ≥ pλ. (VIII.70)

With (VIII.68), we obtain p(λ) = pλ. 11Recall that the discrete spectrum σd(a) consists of all eigenvalues of a.

Page 67: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

VIII.5. THE SPECTRAL THEOREM FOR UNBOUNDED OPERATORS 67

As a corollary, it follows that if σ(a) = σd(a) = λiı∈I , then∫dE ϕ =

∑ı∈I ϕ(i), and

consequently

a =∫dE id =

∑ı∈I

λip(λ). (VIII.71)

As we have seen, this applies in particular to compact operators, but the class of operators withpurely discrete spectrum is considerably larger (e.g. a = 1 is not compact when H is infinite-dimensional, yet has σ(a) = σd(a) = 1).

VIII.5 The spectral theorem for unbounded operators

Formally, the statement of the spectral theorem for possibly unbounded operators is the same asin the bounded case (cf. Theorem VIII.7):

Theorem VIII.12 Let a be a self-adjoint operator on H. Then there exists a unique spectraldensity E on H with the property

a =∫dE(t) t. (VIII.72)

Let us note, however, that this equality entails more than in the bounded case, since the definitionof an operator includes the specification of its domain. By definition, the domain of the right-handside is given by (VIII.40) with ϕ = id; hence Theorem VIII.7 includes the claim that

D(a) = f ∈ H |∫dρf (t) t2 <∞. (VIII.73)

We are going to prove Theorem VIII.12 using a method due to von Neumann [19], which derivesthe spectral theorem for unbounded self-adjoint operators from the spectral theorem for unitaryoperators:

Theorem VIII.13 Let u be a unitary operator on H. Then there exists a unique spectral densityE on H with the properties E(0) = 0 (hence E(t) = 0 for all t ≤ 0), E(2π) = 1 (hence E(t) = 1for all t ≥ 2π), and

u =∫ 2π

0

dE(t) eit. (VIII.74)

The proof is practically the same as for self-adjoint operators, with the following difference. Inthe self-adjoint case we started with polynomials p in a real variable t, defined p(a) in the obviousway, and subsequently defined ϕ(a) for arbitrary bounded (Borel) functions ϕ by approximatingthe latter by polynomials. The properties p(a)∗ = p(a) and p(a) ≤ q(a) if p(t) ≤ q(t) for all(relevant) t then implied the corresponding properties for arbitrary ϕ (see Proposition VIII.10).These properties hold by virtue of the equality a∗ = a. Now, for unitary operators u one insteadhas

u∗u = uu∗ = 1. (VIII.75)

Reflecting this change, we now start from polynomials in the variable exp(it), so that the mapϕ 7→ ϕ(u) is initially defined for functions of the form ϕ(t) =

∑k ck exp(itk), which is mapped to

ϕe(u) :=∑k cku

k (whereas for self-adjoint a this ϕ would be mapped into ϕ(a) =∑k ck exp(ika)).

Hence we may assume that 0 < t ≤ 2π. In particular, writing

e(t) := exp(it), (VIII.76)

one hase 7→ ee(u) = u. (VIII.77)

Clearly, (VIII.75) is reflected by the identity ee = ee = 1, but also more generally one has ϕe(u)∗ =ϕe(u) and ϕe(u) ≤ ψe(u) if ϕ(t) ≤ ψ(t) for all t ∈ (0, 2π] for functions ϕ,ψ of the said form;Proposition VIII.10 holds with a replaced by u and ϕ replaced by ϕe.

Page 68: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

68 CHAPTER VIII. SPECTRAL THEORY FOR SELF-ADJOINT OPERATORS

From this point onwards, the proof of the spectral theorem is the same as in the self-adjointcase. Any positive bounded Borel function on the interval (0, 2π] can be approximated pointwiseby a monotone increasing sequence of polynomials in exp(it), just as it can be thus approximatedby plynomials in t (as we did in the self-adjoint case).12 Proposition VIII.10 holds with a replacedby u and Ia replaced by (0, 2π]. As in (VIII.56), we define the spectral density E by

E(t) = 0 (t ≤ 0);E(t) = (χ(0,t])e(u) (t ∈ (0, 2π]);E(t) = 1 (t > 2π). (VIII.78)

This guarantees the property p(B) = (χB)e(u) (cf. (VIII.57)), from which ϕ(E) = ϕe(u) for simplefunctions ϕ. Approximating a general bounded Borel function ϕ by simple functions ϕn as inthe self-adjoint case, the numerical convergence ϕn → ϕ (pointwise) implies the (strong) operatorconvergence ϕn(E) → ϕ(E) and (ϕn)e(u) → ϕe(u). Since ϕn(E) = (ϕn)e(u), this gives

ϕ(E) = ϕe(u). (VIII.79)

Taking ϕ = e and using (VIII.76) and (VIII.77)(as well as the notation (VIII.39)), we obtain(VIII.74).

To prove Theorem VIII.12 from Theorem VIII.13, von Neumann introduced an operator versionof the Cayley transform. For numbers, this is a map κ : R → C defined by

κ(t) :=t− i

t+ i. (VIII.80)

This maps R homeomorphically (that is, continuously and bijectively with continuous inverse) toT\1 (i.e., the unit circle in the complex plane minus the point 1). Moreover, since formallyκ(−∞) = κ(∞) = 1, it maps the one-point compactification R homeomorphically to T. Theinverse is

κ−1(z) = i1 + z

1− z, (VIII.81)

where z ∈ T\1. Now let a be a self-adjoint operator. We see from Theorem VI.16 that −i ∈ ρ(a),so that a + i : D(a) → H is a bijection with bounded inverse (a + i)−1 : H → D(a). Composingwith a− i : D(a) → H, which is a bijection, too, we see that

κ(a) :=a− i

a+ i= (a− i)(a+ i)−1 (VIII.82)

is a bijection from H to H. In a diagram, we have

H(a+i)−1

−→ D(a) a−i−→ H.

Furthermore, since a∗ = a one computes ‖(a + i)f‖ = ‖(a − i)f‖ for all f ∈ D(a), and puttingf = (a+ i)−1g we find ‖g‖ = ‖κ(a)‖. In other words, κ(a) is an isometry, and since an invertibleisometry is unitary, we conclude that κ(a) is unitary. We note that

R(1− κ(a)) = D(a), (VIII.83)

since 1− κ(a) = 2i(a+ i)−1, and as we have seen (a+ i)−1 : H → D(a) is a bijection.Conversely, let u be a unitary operator with R(1 − u)− = H. (For example, u = 1 does not

satisfy this assumption!) It is an easy exercise to show that this implies that 1− u is injective, sothat 1 − u : H → R(1 − u) is invertible. We may then define an operator on the dense domainD(κ−1(u)) := R(1− u) by

κ−1(u) := i1 + u

1− u= i(1 + u)(1− u)−1. (VIII.84)

12This is a property of Fourier series that we will not prove in this notes.

Page 69: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

VIII.5. THE SPECTRAL THEOREM FOR UNBOUNDED OPERATORS 69

If u = κ(a) then clearly κ−1(u) = a (including its domain), justifying the notation κ−1. Moregenerally, one can show (cf. the next section) that κ−1(u) is self-adjoint; this is not needed for thepresent section. Consequently:

Proposition VIII.14 The operator Cayley transform a 7→ κ(a) establishes a bijection betweenthe class of self-adjoint operators a and the class of unitary operators u with R(1−u)− = H. Thisbijection preserves the spectrum in the sense that

σ(κ(a)) = κ(σ(a)); (VIII.85)σd(κ(a)) = κ(σd(a)). (VIII.86)

We leave the proof of the second part to the reader; it implies that Theorem VIII.11 holds forunbounded operators as well.

Now let u = κ(a) and putϕ(t) = κ−1(exp(it)) (VIII.87)

in (VIII.79); this yields ϕe(u) = κ−1(u) = a, so that

a = i

∫ 2π

0

dE(t)1 + eit

1− eit. (VIII.88)

A chnage of variables t 7→ λ, where λ(t) := κ−1(exp(it)) gives

a =∫

RdE(λ)λ, (VIII.89)

where E(λ) := E(t(λ)), where t(λ) is the unique t ∈ (0, 2π] for which κ(λ) = exp(it). Witha change of notation, this is precisely (VIII.72), up to the domain. According to (VIII.40), thedomain of the right-hand side of (VIII.88) is

Dϕ(E) =

f ∈ H |

∫ 2π

0

d(f,E(t)f)∣∣∣∣1 + eit

1− eit

∣∣∣∣2 <∞

. (VIII.90)

If f ∈ R(1 − u), so that f = (1 − u)g for some g ∈ H, then, using (VIII.42) with ϕ given by(VIII.87),

‖ϕ‖L2((0,2π],ρf ) = ‖ϕ(E)f‖ = ‖(1 + u)g‖ <∞,

since (1 − u)ϕ(E) = (1 + u) by Proposition VIII.10 (for u), while (1 + u) is bounded. Hencef ∈ Dϕ(E) by (VIII.40), so that

R(1− u) ⊆ Dϕ(E). (VIII.91)

Conversely, if f ∈ Dϕ(E) then the integral in (VIII.90) is finite; but∫ 2π

0

d(f,E(t)f)∣∣∣∣1 + eit

1− eit

∣∣∣∣2 = −1 + 4∫ 2π

0

d(f,E(t)f)∣∣∣∣ 11− eit

∣∣∣∣2 , (VIII.92)

since ∣∣∣∣1 + eit

1− eit

∣∣∣∣2 =cos2(t/2)sin2(t/2)

= −1 +1

sin2(t/2)

and 1/ sin2(t/2) = 4/|1 − eit|2, along with dρf = 1. Hence the integral on the right-hand side of(VIII.92) is finite; but ∫ 2π

0

d(f,E(t)f)∣∣∣∣ 11− eit

∣∣∣∣2 = ‖ψ‖2L2((0,2π],ρf )

for ψ : t 7→ 1− exp(it). Hence f ∈ Dψ(E) by (VIII.40), but by (VIII.79) one has ψ(E) = (1+u)−1.It follows that f ∈ D((1 + u)−1) = R(1 + u). It follows that

Dϕ(E) ⊆ R(1− u), (VIII.93)

so with u = κ(a) in (VIII.83) and (VIII.87) we finally obtain

D(a) = R(1− κ(a)) = Dκ−1e,

which is precisely the domain of definition of the right-hand side of (VIII.88).

Page 70: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

70 CHAPTER VIII. SPECTRAL THEORY FOR SELF-ADJOINT OPERATORS

VIII.6 Self-adjoint extensions of symmetric operators

We have defined the Cayley transform for self-adjoint operators (a = a∗), leading to a bijectivecorrespondence between such operators and unitary operators u for which R(1−u)− = H. In thissection we extend the Cayley transform to the class of closed symmetric operators (a ⊂ a∗), whichturn out to be mapped into isometries rather than unitaries. This extension will enable us to solvethe following problem:

Can a symmetric operator be extended to a self-adjoint one, and if so, in which ways?

To illustrate this problem, let us return to Proposition VI.9. The operator a = −id/dx :D(a) → H = L2([0, 1]) is symmetric but not self-adjoint on the domain

D(a) = f ∈ AC([0, 1]) | f ′ ∈ L2 & f(0) = f(1) = 0. (VIII.94)

However, the operator b = −id/dx is self-adjoint on the domain

D(b) = f ∈ AC([0, 1]) | f ′ ∈ L2 & f(0) = f(1). (VIII.95)

We say that b is a self-adjoint extension of a; more generally, this terminology applies to any closedoperator a ⊂ b ⊂ a∗ provided that b∗ = b. We will now develop the machinery to completely solvethe above question, recovering the above example as a special case.

The new situation compared to the self-adjoint case is that for a symmetric but non self-adjointoperator a on a Hilbert space H one no longer has the property R(a± i) = H; indeed, by TheoremVI.16 this equality is equivalent to the self-adjointness of a. However, if a is closed (which we shallalways assume in what follows),13 then the spaces R(a± i) are closed (easy exercise). Furthermore,Lemma VI.12 implies that a ± i is injective, so that in particular a + i : D(a) → R(a + i) is abijection. We may therefore define the Cayley transform (VIII.82) as a map from

D(κ(a)) = R(a+ i) (VIII.96)

to H, with rangeR(κ(a)) = R(a− i). (VIII.97)

The same calculation as in the self-adjoint case then shows that κ(a) is isometric on its domain, andtherefore unitary as a map from the Hilbert space D(κ(a)) ⊂ H to the Hilbert space R(κ(a)) ⊂ H.Such a map is called an isometry. Let us note that the domain is not dense in H (except whena is self-adjoint). One may define an operator on all of H by simple putting κ(a)f = 0 for allf ∈ R(a+ i)⊥ and extension by linearity; in that case one speaks of a partial isometry. Whetherone does this or not is immaterial for what follows. For later use, we define the deficiency indicesof a as

n±(a) := dim(R(a± i)⊥) = dim(N(a∗ ∓ i)); (VIII.98)

cf. Lemma VI.11. The second equality is usually the way to compute the deficiency indices.Repeating the arguments for the self-adjoint case leads to the fundamental result on the Cayleytransform:

Theorem VIII.15 1. The Cayley transform a 7→ κ(a) := (a − i)(a + i)−1 with inverse v 7→κ−1(v) := i(1 + v)(1− v)−1 establishes a bijective correspondence between the class of closedsymmetric operators a on a Hilbert space H and the class of (partial) isometries v for whichR(1− v)− = H.

2. The operator a is self-adjoint iff κ(a) is unitary, and the operator v is unitary iff κ−1(v) isself-adjoint.

13Recall from the text following Definition V.8 that a symmetric operator a is always closable, with closurea− = a∗∗.

Page 71: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

VIII.6. SELF-ADJOINT EXTENSIONS OF SYMMETRIC OPERATORS 71

The proof of these claims is left to the reader as an exercise. It is clear that a is self-adjoint iffn+(a) = n−(a) = 0. Suppose this is not the case. If a has a self-adjoint extension b, then κ(b) is anextension of κ(a), i.e., it coincides with κ(a) on D(κ(a)) ⊂ D(κ(b)) = H, and, being unitary on allof H, κ(b) must be unitary as a map from R(a+ i)⊥ to R(a− i)⊥.14 Conversely, if there exists nosuch unitary map from R(a+ i)⊥ to R(a− i)⊥, then a cannot have a self-adjoint extension. Sincetwo Hilbert spaces are unitarily isomorphic iff they have the same dimension, we have proved:

Proposition VIII.16 A closed symmetric operator a has a self-adjoint extension iff n+(a) =n−(a).

This includes the possibility that both deficiency indices are infinite.15

Suppose one indeed has n+(a) = n−(a). Then any unitary map U : R(a + i)⊥ → R(a − i)⊥

defines a self-adjoint extension aU as follows. First, we define a unitary κ(a)U : H → H by

κ(a)U (f + g) := κ(a)f − Ug, (VIII.99)

where f ∈ D(κ(a)) = R(a+ i) and g ∈ R(a+ i)⊥. Second, we put

aU := κ−1(κ(a)U ). (VIII.100)

Since κ(a)U is unitary, aU is self-adjoint by Theorem VIII.15.16 From (VIII.83) we have

D(aU ) = R(1− κ(a)U ) = R(1− κ(a))⊕R(1 + U) = f + g + Ug | f ∈ D(a), g ∈ R(a+ i)⊥.

By (VIII.100) and (VIII.84), we have, using the fact that g + Ug = (1− κ(a)U )g,

aU (f + g +Ug) = i(1 + κ(a)U )(1− κ(a)U )−1(f + g +Ug) = af + i(1 + κ(a)U )g = af + i(1−U)g.

Conversely, a self-adjoint extension b of a is necessarily of the form b = aU for some unitary U asabove; this follows because κ(b) is an extension of κ(a). Hence we conclude:

Theorem VIII.17 Let a be a closed symmetric operator with n+(a) = n−(a). The self-adjointextensions aU of a are in bijective correspondence with unitary operators U : R(a+i)⊥ → R(a−i)⊥.The domain of aU is

D(aU ) = f + g + Ug | f ∈ D(a), g ∈ R(a+ i)⊥, (VIII.101)

and the action of aU isaU (f + g + Ug) = af + i(1− U)g. (VIII.102)

In the example at the beginning of this section one has a∗ = −id/dx on the domain

D(a∗) = f ∈ AC([0, 1]) | f ′ ∈ L2. (VIII.103)

We compute the deficiency indices using (VIII.98). The space N(a∗ + i) consists of the solutionsf ∈ D(a∗) of a∗f = −if , or df/dx = f . Iteration gives f ∈ C∞([0, 1]), so that we can solvethis equation naively to find f(x) = c exp(x). Similarly for N(a − i). Hence n+ = n− = 1, sothat a indeed has self-adjoint extensions. Since R(a + i)⊥ ∼= R(a − i)⊥ ∼= C, these extensions arelabeled by unitary maps U : C → C. Such a map is given Uf = zf for z ∈ T. A calculation using(VIII.101) then gives

D(az) = f ∈ AC([0, 1]) | f ′ ∈ L2 & f(0) = eiθf(1), (VIII.104)

with exp(iθ) = (1 + ze)/(z + e). Physicists would say that the self-adjoint extensions of a areclassified by a “θ-angle”.

14For the only wat to extend an isometry v : K → H with K ⊂ H to a unitary v : H → H is to defined a unitarymap U : K⊥ → R(v)⊥ and define v by v(f + g) = vf ± ug for f ∈ K and g ∈ K⊥. For obscure historical reasonsone picks the minus sign here.

15In the non-separable case they must then have the same cardinality; in the separable case this is automatic.16Of course, if R(1− κ(a)) is already dense, then certainly R(1− κ(a)U ) is, as κ(a)U is an extension of κ(a).

Page 72: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

72 CHAPTER VIII. SPECTRAL THEORY FOR SELF-ADJOINT OPERATORS

Page 73: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

Chapter IX

Quantum logic

IX.1 Introduction

One of the most spectacular consequences of quantum mechanics, especially in its Hilbert spaceformalism, has been development of a new kind of logic, called quantum logic. This modificationof classical logic was first proposed by Birkhoff and von Neumann in 1936 [4], and is still thesubject of debate and controversy. For none of the traditional views on logic, for example:

• Logic provides the rules of correct thinking (Aristotle);

• Logic gives fundamental truths about the Universe (early Russell);

• Logic gives tautologies (Wittgenstein);

• Logic is a free creation of the human mind (Brouwer);

• Logic is a particular mathematical structure (the modern view),

suggests that logic should depend on empirical discoveries.1 For this reason, quantum logic isusually interpreted from the last point of view, with a footnote adding that it was once suggestedby the mathematical formalism of quantum mechanics. Indeed, most authors agree that evenquantum mechanics has to be understood using classical logic. However, following Mittelstaedt[18], we believe that the rules of correct thinking are indeed described by quantum logic,2, but thatthese rules were only discovered after it became clear from quantum mechanics that classical logicis based on a tacit assumption that may be false. This assumption states that any two propositionssimultaneously have a sharp truth-value (i.e., they are either both true, or both false, or one istrue and the other false). It is precisely this assumption that is denied by quantum mechanics.As we shall see, in classical logic the proposition (x and y) or (x and not-y) is equal to x. Thisis no longer the case in quantum logic, because x and y may not simultaneously have a sharptruth-value. This gives quantum logic a probabilistic character from the outset.

IX.2 Propositional logic

We will only consider the simplest kind of logical structure in this chapter, called propositionallogic.3 This means that one deals with a structure specified by:

1. Variables;

2. Logical connectives, such as equal to, and, or, not, and implies;1Even in Russell’s view logic precedes natural science.2And they do so precisely because they tell us when two propositions are equivalent, consistent with Wittgen-

stein’s view.3See, for example, [27] for a quick introduction.

73

Page 74: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

74 CHAPTER IX. QUANTUM LOGIC

3. Grammatical rules;

4. Deductive rules;

5. Valuations.

The variables x, y, . . . are interpreted as elementary propositions, which, using the connectivesmay be combined into general proposition. A proposition is a statement to which some truth-valuemay be assigned. For example, the statement “the energy” is not a proposition. However, thestatement “the energy lies between a and b” is a proposition (once the system to which it applieshas been specified). Hence a physical observable does not necessarily define a logical proposition,but, as we shall see, the converse is true.

We will use the following symbols for the logical connectives:

• x ∨ y means x or y;

• x ∧ y means x and y;

• x∼ means not-x;

• x⇒ y means x implies y (as a proposition);

• x→ y means x implies y (as a variable);

• x ≡ y means x is equivalent to y.4

The subtle difference between implication ⇒ and “material implication” →, introduced by Russelland Whitehead, is that x → y is again a variable, whereas x ⇒ y is a statement about vari-ables. This distinction is sometimes useful.5 As far as the truth of propositions is concerned, theconnectives ⇒ and → have to be such that x⇒ y is true iff x→ y is true.

Other symbols may occur as well: we will use

• 1 for the tautological proposition (that is always true);

• o for the contradictory proposition (that is always false).

The connectives are typically not all independent; they are related by the deductive rules. Forexample, in classical logic one may take the symbols → and ∼ as basic, and define the others by:

1. x ∨ y := (x→ y) → x

2. x ∧ y := (x∼ ∨ y∼)∼ (De Morgan’s rule).

Furthermore, x ≡ y means that x⇒ y and y ⇒ x both hold.If the symbols ∨, ∧, and ≡ are primitive, these would be deductive rules. Other deductive rules

depend on the specific logic. In both classical and quantum logic the following rules hold:6

1. 0 → x (Law of absurdity);7

2. x∼∼ = x (Law of negation);

3. x ∨ x∼ = 1 (Law of excluded middle third);

4. x ∧ x∼ = o (Law of contradiction);

5. (x ∨ y)∼ = x∼ ∧ y∼ (De Morgan’s rule);4This means that x and y always have the same values in any valuation, see below.5In classical logic x→ y is given by x∼ ∨ y.6These rules are not all independent.7The fact that a contradiction implies any proposition is widely felt as a weakness of both classical and quantum

logic. Russell, who defended this law and who was a convinced atheist, was once asked how 2 + 2 = 5 implies thatRussell was the Pope. He answered that 2 + 2 = 5 implies 1 = 2, which proved that he and the Pope (namely theright-hand-side), were really one (namely the left-hand side). Hence Russell was the Pope.

Page 75: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

IX.3. THE LATTICE OF A PROPOSITIONAL LOGIC 75

6. (x ∧ y)∼ = x∼ ∨ y∼ (De Morgan’s rule);

7. (x ∧ (x→ y)) → y (Modus ponens);

8. (y∼ ∧ (x→ y)) → x∼ (Modus tollens);

9. (y ∧ (x∼ → y∼)) → x (Reductio ad absurdum).

These deductive rule specify tautologies in the logic under consideration, in the sense that thecorresponding propositions are true independently of the truth of the variables occurring therein.Similarly, contradictions are propositions that are always false in this sense.

As we have already mentioned, the principal distinction between classical logic and quantumlogic is that in the former the formula

x = (x ∧ y) ∨ (x ∧ y∼) (IX.1)

is a tautology (see below), whereas in the latter it may be false (depending on x and y).Truth values are assigned to propositions via a so-called valuation. In classical logic a valuation

is a map τ : V → 0, 1, where V = x, y, z, . . . is the set of variables, such that τ(x) = 0 meansthat x is false, and τ(x) = 1 means that x is true. The truth of general propositions may then bedetermined from the truth of the variables occurring in it via truth tables.8 In classical logic alltruth tables are determined by

1. τ(x∼) = 0 when τ(x) = 1 and vice versa;

2. τ(x→ y) = 0 when τ(x) = 1 and τ(y) = 0, and τ(x→ y) = 1 otherwise.

For example, defining ∨ and ∧ in terms of ∼ and →, one finds that τ(x∨ y) = 1 when τ(x) = 1 orτ(y) = 1, and τ(x ∨ y) = 0 when τ(x) = 0 and τ(y) = 0. Also, one computes that (IX.1) is alwaystrue.

Analogously to the discussion of pure and mixed states in Section VII.5, one could regard avaluation of the above type as “pure”, and also consider more general valuations as maps τ : V →[0, 1]. We will demonstrate this possibility below both in classical and quantum logic.

IX.3 The lattice of a propositional logic

Through a procedure that goes back to Boole in 1854 and was brought into its modern form byLindenbaum and Tarski, one may extract the essence of a given propositional logic L by associatinga certain algebraic structure L to it, called a lattice.9 The idea is that the elements of the algebraL are equivalence [f ] classes of formulae f in L (where a formula is a well-formed expression in thevariables and the logical connectives ∨, ∧, and ∼), where two formulae are considered equivalentwhen they are logically equivalent.10 Furthermore, the role of implication ⇒ in L is taken over bya partial ordering ≤ in L, in such a way that [f ] ≤ [g] when f ⇒ g in L. It turns out that afterforming equivalence classes the logical meaning of ∨ and ∧ is in fact determined by that of ⇒, andvice versa.

Definition IX.1 A lattice (L,≤) is a partially ordered set (poset)11 in which any two elementsx, y have a supremum (or least upper bound) x ∨ y,12 and an infimum (or greatest lower bound)x ∧ y.13

This definition is equivalent to the following:8What follows could have been expressed in tabular form.9The classical work on this subject is [2].

10That is, related by ≡, which in turn means that they are always (i.e., in any vaulation) simultaneously true orfalse.

11A partial ordering on a set S is a relation (i.e., a subset of S × S) called ≤ that is reflexive in that x ≤ x forall x, antisymmetric in that x ≤ y and y ≤ x imply x = y, and transitive in that x ≤ y and y ≤ z imply x ≤ z.

12That is, x ≤ x ∨ y and y ≤ x ∨ y, and if x ≤ z and y ≤ z for some z, then x ∨ y ≤ z.13I.e., x ≥ x ∧ y and y ≥ x ∧ y, and if x ≥ z and y ≥ z for some z, then x ∧ y ≥ z.

Page 76: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

76 CHAPTER IX. QUANTUM LOGIC

Definition IX.2 A lattice (L,∨,∧) is a set L equipped with two idempotent,14 commutative, andassociative operations ∨,∧ : L × L → L that satisfy x ∨ (y ∧ x) = x and x ∧ (y ∨ x) = x.

Given (L,∨,∧), the partial ordering is defined by x ≤ y if x ∧ y = x.The “classical” example of a lattice is obtained by taking a set S and defining L as the power

set 2S of S (i.e., the set of all subsets of S). The lattice structure of (L,≤) consists of ≤ = ⊂, thecorresponding structure of (L,∨,∧) being ∨ := ∪ and ∧ := ∩.

A lattice L is called complete when every subset of L has a supremum as well as an infimum.We do not assume this to be the case, but we do assume that L itself has a supremum, denoted by1, as well as an infimum, called o. Hence o ≤ x ≤ 1 for all x ∈ L. The above lattice 2S is actuallycomplete, with 1 = S and o = ∅.

In order to reflect the negation ∼ of a logic in its associated lattice, one introduces the notionof orthocomplementation.

Definition IX.3 An orthocomplementation on a lattice (L,≤) with o and 1 is a map x 7→ x∼,satisfying (for all x, y ∈ L)

• x∼∼ = x.

• x ≤ y ⇐⇒ y∼ ≤ x∼.

• x ∧ x∼ = o.

• x ∨ x∼ = 1.

It follows that 1∼ = o and o∼ = 1, and that de Morgan’s laws (x∨y)∼ = x∼∧y∼ and (x∧y)∼ =x∼∨y∼ hold. A lattice with an orthocomplementation is called an orthocomplemented lattice.For example, the lattice L = 2S is orthocomplemented under the set-theoretic complement, i.e.,

A∼ = Ac = S\A. (IX.2)

The upshot of all these definitions is that many interesting logics L, including classical logicand quantum logic, define an orthocomplemented lattice L in the following way: the elements of Lare equivalence [f ] classes of formulae f in L, the partial ordering in L is given by [f ] ≤ [g] whenf ⇒ g in L, the orthocomplementation is given by [f ]∼ = [f∼], and when the whole operation ofpassing from L to L is consistent it should follow that [f ] ∨ [g] = [f ∨ g] and [f ] ∧ [g] = [f ∧ g].

IX.4 The logic of classical physics

A lattice is called distributive when

x ∧ (y ∨ z) = (x ∧ y) ∨ (x ∧ z) (IX.3)

for all x, y, z, as well as the same property with ∨ and ∧ interchanged.15 It can be shown thatthe orthocomplemented lattice defined by a classical logic is distributive, and this is often takenas the defining characteristic of classical logic. A distributive orthocomplemented lattice is calledBoolean.16 As we have seen, the set 2S of subsets of a set S defines a Boolean lattice if one takes≤ = ⊂ and ∼ as set complementation in S. Conversely, using the axiom of choice, Stone provedin the 1930s that every Boolean lattice is isomorphic to the lattice of a certain class of subsets ofa certain topological space.17

Classical physics provides natural examples of this set-theoretic realization of classical logic. Toexplain this, let M be the phase space of a classical physical system. The elementary propositionsabout this system are typically of the form: “the observable f (defined as a Borel function f : M →

14An idempotent map · S × S → S satisfies x · x = x.15These are equivalent when the lattice is orthocomplemented.16Boolean lattices are often called Boolean algebras.17Such a space is extremely disconnected, and the subsets in question are the “clopen” (closed and open) ones.

Page 77: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

IX.5. THE LOGIC OF QUANTUM MECHANICS 77

R) takes some value in B ⊂ R (where B is a Borel set). For reasons to emerge in a second, we callthis proposition χB(f). Any point x in M defined a valuation τx on such elementary propositions,in that τx(χB(f)) = 1 (i.e., χB(f) is true) when f(x) ∈ B, and τx(χB(f)) = 0 (i.e., χB(f) is false)when f(x) /∈ B. Defining χB(f) (which so far was just a name) as a function on M by

χB(f) := χf−1(B), (IX.4)

we see thatτx(χB(f)) = χB(f)(x). (IX.5)

In other words, χB(f) takes the value 1 at those x ∈ M for which the corresponding propositionis true, and equals 0 at those x for which it is false.

More generally, any state P (as defined in physics; see Section VII.5) defines a generalizedvaluation τP by restricting it to elementary propositions of the type χB(f). In other words,recalling that P is a probability measure on M , we put

τP (χB(f)) = P (χB(f)) =∫M

dP (x)χB(f)(x); (IX.6)

cf. (VII.32). Such a generalized valuation takes values in [0, 1], and gives a probabilistic verdict onthe truth of the elementary propositions of classical physics.

Thus we infer that any proposition is equivalent to an observable of the form χA, where A is a(Borel) subset of M . Equivalently, we may just say that the equivalence classes of the propositionsare given by the (Borel) subsets A ⊂M itself.18 The associated lattice operations are

≤ = ⊂; (IX.7)A∼ = Ac; (IX.8)

A ∨B = A ∪B; (IX.9)A ∧B = A ∩B; (IX.10)

1 = M ; (IX.11)o = ∅. (IX.12)

In terms of the observables f of the form χA, these are

f∼ = 1− f ; (IX.13)f ∨ g = max(f, g); (IX.14)f ∧ g = min(f, g); (IX.15)

1 = 1M ; (IX.16)o = 0M . (IX.17)

One may check the distributivity of this lattice, as well as the fundamental tautology (IX.1),which just comes down to the set-theoretic identity (for any A,B ⊂M)

A = (A ∩B) ∪ (A ∩Bc). (IX.18)

IX.5 The logic of quantum mechanics

As we have seen, in quantum mechanics the physical observables are represented by self-adjointoperators on a Hilbert space H. The propositions defined by such an observable a are of the form“(upon measurement) a takes values in B ⊂ R”, just like in classical physics (with the differencethat in quantum theory some measurement context has to be specified). It follows from Theorem

18Elementary propositions of the form “B” or “χB” are sometimes called “yes-no questions”, because they canbe either true (yes) or false (no).

Page 78: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

78 CHAPTER IX. QUANTUM LOGIC

VIII.11 and the standard interpretation of quantum mechanics (see (VII.5), (VIII.30)), and sur-rounding text) that such a proposition is represented by the observable pB defined in (VIII.29),where E is the spectral density defined by a; in other words (cf. (VIII.57)), the proposition inquestion is the projection χB(a). Note that the passage from the verbal proposition above to theoperator χB(a) is entirely analogous to the procedure in classical physics leading to (IX.4).

According to (VII.5) or (VIII.30), a pure state ψ ∈ H of the system defines a probabilisticvaluation τψ (taking values in [0, 1]) on the class of such propositions by means of

τψ(χB(a)) = (ψ, χB(a)ψ) = ‖χB(a)ψ‖2. (IX.19)

This means that χB(a) is possibly true, with probability given by the right-hand side of (IX.19).In particular, χB(a) is certainly true when ψ ∈ χB(a)H, and false when ψ ∈ (χB(a)H)⊥. Thecase of general states will be picked up in the next section.

Consequently, all propositions about a given quantum-mechanical system are equivalent to pro-jections on H. Because of the bijective correspondence between projections p and closed subspacespH of H, equivalence classes of quantum-mechanical propositions may alternatively be said to begiven by closed subspaces K of H.19

Given the standard interpretation of quantum mechanics and the desirability to preserve asmuch as possible of classical logic, there is practically no choice in the definition of the logi-cal connectives relating quantum-mechanical propositions, or equivalently, in choosing the latticeoperations on the set L(H) of all closed subspaces (or projections) on H. Indeed, if we defineimplication by ≤ = ⊂, as in classical logic, it follows that

K ∧ L = K ∩ L; (IX.20)K ∨ L = K ⊕ L := f + g | f ∈ K, g ∈ L−. (IX.21)

It follows that

1 = H; (IX.22)o = 0. (IX.23)

The orthocomplementation is dictated by the probability interpretation: if K is true in a stateψ, then K⊥ is false, and vice versa. Hence we must define

K∼ = K⊥, (IX.24)

The material implication → turns out to be given by

(K → L) = K⊥ ⊕ (K ∩ L), (IX.25)

since only this prescription guarantees that K → L = H (which means that K → L is true) iffK ⊂ L (which means that K ⇒ L is true).

In this way one obtains a lattice L(H), which is supposed the codify the logical structureof quantum-mechanical propositions. The lattice L(H) is the simplest example of a quantumlogic. This logic shares a numer of properties with classical logic: for example, the eight lawslisted in Section IX.2 (starting with the Law of negation) all hold in quantum logic. Nonetheless,quantum logic is not isomorphic to any classical logic, as follows for example from the fact that thedistributive law (IX.3) generally fails to hold. In addition, the equality (IX.1), which in classicallogic is a tautology, does not hold in quantum logic either. In fact, one has

K = (K ∧ L) ∨ (K ∧ L⊥) (IX.26)

in L iff K and L are either mutually orthogonal or one is contained in the other (the correspondingprojections then commute). Thus the quantum logic L(H) contains infinitely many classical logics;

19The bijection between closed subspaces of H and projections on H is reflected in classical physics by thebijection between (Borel) subsets of the phase space M and the characteristic functions of these subsets. As inclassical physics, a projection p defines a yes-no questions, which is reflected by the fact that σ(p) = 0, 1 (exceptwhen p = 1 or p = 0).

Page 79: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

IX.6. GLEASON’S THEOREM 79

each family of mutually commuting projections defines a Boolean subalgebra of L(H), and therebya classical logic. These different classical logics stand in a certain relationship to each other, butthey cannot be combined into a single classical logic.

The failure of (IX.1) for subspaces K and L whose projections do not commute is, of course,mysterious from a classical point of view. The point of view of Birkhoff and von Neumann, whofirst proposed all this [4], was that quantum mechanics forces us to weaken classical logic. In anattempt to characterize quantum logic, they proposed to replace the distributive law (IX.3) by theproperty of modularity; this proposal is now obsolete, since the modular law only holds for finite-dimensional Hilbert space. Instead, most attempts to characterize quantum logic are nowadaysbased on the fact that L(H) satisfies the property of orthomodularity, viz.

x ∧ (x∼ ∨ (x ∧ y)) ≤ y. (IX.27)

This property weakens the distributive law, in the sense that substituting y = x∼ and z = x∧ y in(IX.3) would give x∧(x∼∨(x∧y)) = x∧y, which is indeed ≤ y. The modularity law can be restatedin a number of equivalent ways. For example, an orthocomplemented lattice is orthomodular iffx ≤ y is equivalent to x∼ ∨ (x ∧ y) = 1. This is true in L(H), and justifies the definition (IX.25)of material implication (cf. [18]). However, endless research in this direction has failed to find ana priori or even physical justification of orthomodularity.20

IX.6 Gleason’s theorem

We have already seen that a quantum-mechanical pure state ψ defines a [0, 1]-valued valuation onthe quantum logic L(H) by means of (IX.19). More generally, a density matrix ρ on H defines sucha valuation by simply restricting the pairing (VII.39)) between states and observables to elementarypropositions (i.e., projections). Thus a density matrix ρ defines a map τρ : L(H) → [0, 1] by

τρ(p) = Tr ρp. (IX.28)

For a pure state ρ that is given by the projection onto a unit vector ψ ∈ H, one of course recovers(IX.19).

The map p 7→ τρ(p) satisfies a number of properties reminiscent of probability measures (cf.(VII.29) -(VII.31)), namely (writing τ for τρ):

τ(0) = 0; (IX.29)τ(1) = 1; (IX.30)

τ(∨ipi) =∑i

τ(pi) (IX.31)

for any countable family of mutually orthogonal projections pi. Conversely, Gleason’s theoremstates that any map τ : L(H) → [0, 1] that satisfies these properties is of the form τ = τρ with(IX.28), for some density matrix ρ.21 This provides some justification for the assumption made byphysicists that quantum-mechanical states are density matrices.22

20In our opinion, the best attempt so far is as follows [18]. Let us call two elements x and y of an orthocomple-mented lattice L compatible when (IX.1) holds. Then L is orthomodular iff this relation is symmetric (that is, xis compatible with y when y is compatible with x).

21At least when dim(H) > 2.22A similar justification is provided by von Neumann’s theorem, which states that any linear map ρ : B(H) →

C that satisfies ρ(a∗a) ≥ 0 for all a ∈ B(H), ρ(1) = 1, and (IX.31) is given by (VII.39).

Page 80: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

80 CHAPTER IX. QUANTUM LOGIC

Appendix: bra and ket notation

Dirac introduced a notation for vectors and functionals that is widely used by physicists, butwhich continues to confuse mathematicians. It should be mentioned that Dirac never talked aboutHilbert space and never defined what his linear spaces precisely were. Restricted to Hilbert space,what he meant was this. Vectors in H are denoted |ψ〉, where in these notes we’d simply say ψ.Our inner product (ϕ,ψ) is written as 〈ϕ|ψ〉, with the same properties of being linear in the secondvariable and antilinear in the first. If a is an operator, Dirac wrote 〈ϕ|a|ψ〉 for our (ϕ, aψ). Notealso that Dirac denoted complex conjugation by a ∗, so that his 〈ϕ|ψ〉∗ is the same as our (ϕ,ψ),and adjoints by a dagger, so that his a† is our a∗.

However, some of Dirac’s notation goes beyond the Hilbert space framework. For example, ifH = L2(R) and ψ ∈ H, Dirac wrote 〈x|ψ〉 for ψ(x), and, quite confusingly, 〈p|ψ〉 for ψ(p) (i.e., theFourier transform of ψ). Some notation of this style was later made rigorous by the formalism ofrigged Hilbert spaces (also called Gelfand triples). See, for example, [7].

Page 81: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

Bibliography

[1] E.G. Beltrametti & G. Cassinelli, The logic of quantum mechanics ( Addison-Wesley, Reading,MA, 1981).

[2] G. Birkhoff, Lattice Theory (Third corrected printing, AMS, Providence, 1993).

[3] G. Birkhoff & E. Kreyszig, The establishment of functional analysis, Historia Mathematica11, 258–321 (1984).

[4] G. Birkhoff & J. von Neumann, The logic of quantum mechanics, Ann. Math. 37, 823–843(1936).

[5] A. Connes Noncommutative Geometry (Academic Press, San Diego, 1994).

[6] J. Dieudonne, History of functional analysis (North-Holland, Amsterdam, 1981).

[7] S. J. L. van Eijndhoven & J. de Graaf, A Mathematical Introduction to Dirac’s Formalism(North-Holland Publishing Co., Amsterdam, 1986).

[8] G. Gallavotti, The Elements of Mechanics (Springer, Heidelberg, 1983).

[9] P.R. Halmos, The legend of John von Neumann, The American Mathematical Monthly 80,382-394 (1973).

[10] S.J. Heims, John von Neumann and Norbert Wiener: From Mathematics to the Technologiesof Life and Death (MIT Press, Cambridge, Mass., 1980).

[11] P.D. Hislop & I.M. Sigal, Introduction to Spectral Theory, Springer, New York, 1996.

[12] M. Jammer, The Conceptual Development of Quantum Mechanics (McGraw-Hill, New York,1966).

[13] Klaas Landsman, Requiem voor Newton (Contact, Amsterdam, 2005).

[14] N.P. Landsman, Mathematical Topics Between Classical and Quantum Mechanics (Springer,New York, 1998).

[15] N.P. Landsman, Lecture Notes on C∗-algebras and K-theory,http://remote.science.uva.nl/ npl/CK.html, Part I.

[16] N. Macrae, John von Neumann: The Scientific Genius Who Pioneered the Modern Computer,Game Theory, Nuclear Deterrence, and Much More (Pantheon Books, 1992).

[17] J. Mehra & H. Rechenberg, The Historical Development of Quantum Theory, Vol.6: TheCompletion of Quantum Mechanics 1926-1941 (Springer-Verlag, New York, 1982).

[18] P. Mittelstaedt, Quantum Logic (Reidel, Dordrecht, 1987).

[19] J. von Neumann, Mathematische Grundlagen der Quantenmechanik (Springer, Berlin, 1932).

[20] G.K. Pedersen, Analysis Now (Springer, New York, 1989 and 1995).

81

Page 82: Lecture Notes on Hilbert Spaces and Quantum Mechanicslandsman/HSQM.pdf · VIIQuantum mechanics and Hilbert space 45 ... I.1 Introduction The concept of a Hilbert space is seemingly

82 BIBLIOGRAPHY

[21] J. Pier, Mathematical analysis during the 20th century (Oxford University Press, New York,2001).

[22] M. Reed and B. Simon, Methods of Modern Mathematical Physics, Vol. I: Functional Analysis(Academic Press, San Diego, 1980).

[23] R. Siegmund-Schultze, The origins of functional analysis. H. N. Jahnke (ed.), A History ofAnalysis, pp. 385–408 (AMS, Providence, 2003).

[24] F. Riesz and B. Sz.-Nagy, Functional Analysis (F. Ungar, New York, 1955).

[25] S.M. Ulam et al., John von Neumann 1903-1957, Bulletin of the American MathematicalSociety 64 (3), 1958.

[26] J. Weidmann, Lineare Operatoren in Hilbertraumen, Vols. I&II (Teubner, Stuttgart, 2000&2003).

[27] R. Wallace Garden, Modern Logic and Quantum Mechanics (Adam Hilger, Bristol, 1984).