Lnotes Mathematical Found QM Temp
Lnotes Mathematical Found QM Temp
Stephan Fackler
Version: July 17, 2015
Contents
Introduction iii
3 Distributions 103
3.1 The Space of Distributions . . . . . . . . . . . . . . . . . . . . 103
3.2 Tempered Distributions . . . . . . . . . . . . . . . . . . . . . . 109
3.2.1 The Fourier Transform of Tempered Distributions . . 111
3.3 The Nuclear Spectral Theorem . . . . . . . . . . . . . . . . . . 116
i
Contents
Bibliography 129
ii
Introduction
These lecture notes were created as a companion to the lecture series hold
together with Kedar Ranade in the summer term 2015 under the same title.
The lecture was aimed at both master students of physics and mathematics.
Therefore we required no prior exposure to neither the apparatus of func-
tional analysis nor to quantum physics. The mathematical background was
presented in my lectures, whereas the students were introduced to the physics
of quantum mechanics in Kedar’s part of the lecture.
The aim of the lectures was to present most of the mathematical results and
concepts used in an introductory course in quantum mechanics in a rigorous
way. Since physics students usually have no background in Lebesgue integra-
tion, a short primer on this topic without proofs is contained in the first chap-
ter. Thereafter the fundamentals of the theory of Hilbert and Sobolev spaces
and their connection with the Fourier transform are developed from scratch.
It follows a detailed study of self-adjoint operators and the self-adjointness of
important quantum mechanical observables, such as the Hamiltonian of the
hydrogen atom, is shown. Further, the notes contain a careful presentation
of the spectral theorem for unbounded self-adjoint operators and a proof
of Stone’s theorem on unitary groups which is central for the description of
the time evolution of quantum mechanical systems. The spectral theory of
self-adjoint operators and Hamiltonians is only covered in a very rudimentary
manner.
In the last part a short introduction to the theory of distributions is given.
Further, we present the nuclear spectral theorem which gives the spectral
decomposition of self-adjoint operators in a form very natural for physicists.
The appendix covers precise mathematical statements of the postulates of
quantum mechanics presented in the course for further easy reference.
iii
A Crash Course in Measure Theory 1
In classical quantum mechanics (pure) a quantum mechanical system is de-
scribed by some complex Hilbert space. For example, the (pure) states of a
single one-dimensional particle can be described by elements in the Hilbert
space L2 (R) as introduced in introductory courses in quantum mechanics. A
natural first attempt to mathematically define this space is the following:
( Z∞ )
2 2
L (R) = f : R → C : f|[−n,n] Riemann-int. for n ∈ N and |f (x)| dx < ∞ .
−∞
However, there are several issues. First of all, the natural choice
Z∞ !1/2
2
kf k2 B |f (x)| dx
−∞
does not define a norm on L2 (R) as there exist functions 0 , f ∈ L2 (R) with
kf k2 = 0. This problem can easily be solved by identifying two functions
f , g ∈ L2 (R) whenever kf − gk2 = 0. A more fundamental problem is that the
above defined space is not complete, i.e. there exist Cauchy sequences in L2 (R)
which do not converge in L2 (R). Therefore one has to replace L2 (R) as defined
above by its completion. This is perfectly legitimate from a mathematical
point of view. However, this approach has a severe shortcoming: we do not
have an explicit description of the elements in the completion. Even worse,
we do not even know whether these elements can be represented as functions.
To overcome these issues, we now introduce an alternative way to integra-
tion, finally replacing the Riemann-integral by the so-called Lebesgue-integral.
In order to be able to introduce the Lebesgue-integral we need first a rigorous
method to measure the volume of subsets of Rn or more abstract sets which
then can be used to define the Lebesgue integral.
The material covered in this chapter essentially corresponds to the basic
definitions and results presented in an introductory course to measure theory.
We just give the definitions with some basic examples to illustrate the concepts
and then state the main theorems without proofs. More details and the proofs
can be learned in any course on measure theory or from the many excellent
text books, for example [Bar95] or [Rud87]. For further details we guide the
interested reader to the monographs [Bog07].
1
1. A Crash Course in Measure Theory
A = A1 ∪ . . . ∪ An and B = B1 ∪ . . . ∪ Bn
Using such paradoxical decompositions we see that m must agree for all
bounded subsets of Rn with non-empty interiors. For example, by splitting a
cube Q into two smaller parts, we see that m(Q) ∈ (0, ∞) leads to a contradic-
tion. Hence, it is impossible to measure the volume of arbitrary subsets of Rn
in a reasonable way!
Remark 1.1.2. Of course, we all know that in physical reality such a paradox
does not occur. Indeed, the decompositions given by the Banach–Tarski
paradox are not constructive and therefore cannot be realized in the real
world. More precisely in mathematical terms, the proof of the Banach–Tarski
paradox requires some form of the axiom of choice.
(a) ∅ ∈ Σ,
The tuple (Ω, Σ) is called a measurable space and the elements of Σ are called
measurable.
2
1.1. Measure Spaces
Note that it follows from the definition that for A, B ∈ Σ one also has
A ∩ B ∈ Σ and B \ A ∈ Σ. The closedness of Σ under countable unions may
be the less intuitive of the above defining properties. It guarantees that
σ -algebras behave well under limiting processes which lie at the hearth of
analysis. We now give some elementary examples of σ -algebras.
Example 1.1.4. (i) Let Ω be an arbitrary set. Then the power set P (Ω) is
clearly a σ -algebra.
is called the Borel σ -algebra on Ω. One can show that B(Rn ) is the smallest
σ -algebra that is generated by elements of the form [a1 , b1 ) × · · · [an , bn ) for
ai < bi , i.e. by products of half-open intervals.
3
1. A Crash Course in Measure Theory
Proposition 1.1.7. Let Ω1 and Ω2 be two normed vector spaces or more generally
metric or topological spaces. Then every continuous mapping f : Ω1 → Ω2 is
measurable. Further, every monotone function f : R → R is measurable.
(i) µ(∅) = 0.
P∞
(ii) µ(∪n∈N An ) = n=1 µ(An ) for all pairwise disjoint (An )n∈N ⊂ Σ.
The triple (Ω, Σ, µ) is a measure space. If µ(Ω) < ∞, then (Ω, Σ, µ) is called a
finite measure space. If µ(Ω) = 1, one says that (Ω, Σ, µ) is a probability space.
One can deduce from the above definition that a measure satisfies µ(A) ≤
µ(B) for all measurable A ⊂ B and µ(∪n∈N Bn ) ≤ ∞
P
n=1 µ(Bn ) for arbitrary
(Bn )n∈N ⊂ Σ. Moreover, one has µ(A \ B) = µ(A) − µ(B) for measurable B ⊂ A
whenever µ(B) < ∞. We begin with some elementary examples of measure
spaces.
Example 1.1.10. (i) Consider (Ω, P (Ω)) for an arbitrary set Ω and define
µ(A) as the number of elements in A whenever A is a finite subset and
µ(A) = ∞ otherwise. Then µ is a measure on (Ω, P (Ω)).
4
1.2. The Lebesgue Integral
δa : P (Ω) → R≥0
1 if a ∈ A
A 7→ .
0 else
for all products with ai < bi . The measure λ is called the Lebesgue measure on
Rn .
Of course, one can also restrict the Lebesuge measure to (Ω, B(Ω)) for
subsets Ω ⊂ Rn . The uniqueness in the above theorem is not trivial, but
essentially follows from the fact that the products of half-open intervals used
in the above definition generate the Borel-σ -algebra and are closed under finite
intersections. The existence is usually proved via Carathéodory’s extension
theorem.
(i) For a simple function f : Ω → R≥0 given by f = nk=1 ak 1Ak as above one
P
5
1. A Crash Course in Measure Theory
6
1.2. The Lebesgue Integral
One can show that the set C of all x ∈ Ω for which the above limit exists
is measurable. It follows easily from this fact the function f : Ω → C is
measurable as well. Note further that because of C ⊂ N one has µ(C) = 0.
Hence, the Lebesgue integral of f is independent of the concrete choice of
the values at the non-convergent points and therefore the choice does not
matter for almost all considerations. We make the agreement that we will
always define the pointwise limit of measurable functions in the above way
whenever the limit exists almost everywhere. This is particularly useful for the
formulation of the following convergence theorems for the Lebesgue integral.
7
1. A Crash Course in Measure Theory
Note that the monotonicity assumption is crucial for the theorem. In fact,
in general one cannot switch the order of limits and integrals as the following
example shows.
Z Z
lim 1[n,n+1] dλ = 1 , 0 = lim 1[n,n+1] dλ.
n→∞ R Ω n→∞
For example, (N, P (N)) together with the counting measure or the measure
spaces (Rn , B(Rn ), λ) for n ∈ N, where λ denotes the Lebesgue measure, are
σ -finite. Moreover, every finite measure space and a fortiori every probability
space is σ -finite. For an example of a non-σ -finite measure space consider
(R, P (R)) with the counting measure.
8
1.2. The Lebesgue Integral
For example, one has B(Rn )⊗B(Rm ) = B(Rn+m ) which can be easily verified
using the fact that products of half-open intervals generate B(Rn ). It follows
from the characterizing property of the Lebesgue measure λn on (Rn , B(Rn ))
that for all n, m ∈ N the measure λn+m is a product measure of λn and λm . One
can show that there always exists a product measure for two arbitrary measure
spaces. In most concrete situations there exists a uniquely determined product
measure as the following theorem shows.
Theorem 1.2.7. Let (Ω1 , Σ1 , µ1 ) and (Ω2 , Σ2 , µ2 ) be two σ -finite measure spaces.
Then there exists a unique product measure on (Ω1 ×Ω2 , Σ1 ⊗Σ2 ) which is denoted
by µ1 ⊗ µ2 .
is finite, then one has for the product and iterated integrals
Z Z Z
f (x, y) d(µ1 ⊗ µ2 )(x, y) = f (x, y) dµ2 (y) dµ1 (x)
Ω1 ×Ω2 Ω1 Ω2
Z Z
= f (x, y) dµ1 (x) dµ2 (y).
Ω2 Ω1
Note that there are also variants of Fubini’s theorem (not in the above gen-
erality) for non σ -finite measure spaces. However, this case is more technical
and rarely used in practice and therefore we omit it.
9
1. A Crash Course in Measure Theory
Definition 1.3.1 (Lp -spaces). Let (Ω, Σ, µ) be a measure space. For p ∈ [1, ∞)
we set
( Z )
p p
L (Ω, Σ, µ) B f : Ω → K measurable : |f | dµ < ∞ ,
Ω
Z !1/p
p
kf kp B |f | dµ .
Ω
For p = ∞ we set
Note that the space L1 (Ω, Σ, µ) agrees with the space L1 (Ω, Σ, µ) previ-
ously defined in Definition 1.2.1. One can show that (Lp (Ω, Σ, µ), k·kp ) is a
semi-normed vector space, i.e. k·kp satisfies all axioms of a norm except for def-
initeness. Here, the validity of the triangle inequality, the so-called Minkowski
inequality, is a non-trivial fact. If one identifies two functions whenever they
agree almost everywhere, one obtains a normed space.
Definition 1.3.2 (Lp -spaces). Let (Ω, Σ, µ) be a measure space and p ∈ [1, ∞].
The space Lp (Ω, Σ, µ) is defined as the space Lp (Ω, Σ, µ) with the additional
agreement that two functions f , g : Ω → K are identified with each other
whenever f − g = 0 almost everywhere.
Recall that a normed vector space or more generally a metric space is called
complete if every Cauchy sequence converges to an element in the space. A
sequence (xn )n∈N in a normed vector space (V , k·k) is called a Cauchy sequence
if for all ε > 0 there exists n0 ∈ N such that kxn − xm k ≤ ε for all n, m ≥ n0 .
Using this terminology we have
Theorem 1.3.4 (Riesz–Fischer). Let (Ω, Σ, µ) be a measure space and p ∈ [1, ∞].
Then Lp (Ω, Σ, µ) is a Banach space.
10
1.3. Lebesgue Spaces
Proposition 1.3.5. Let (Ω, Σ, µ) be a measure space and p ∈ [1, ∞). Further
suppose that fn → f in Lp (Ω, Σ, µ). Then there exist a subsequence (fnk )k∈N and
g ∈ Lp (Ω, Σ, µ) such that
Proposition 1.3.6. Let Ω ⊂ Rn be open and p ∈ [1, ∞). Then Cc (Ω), the space of
all continuous functions on Ω with compact support (in Ω), is a dense subspace of
Lp (Ω).
Proposition 1.3.8. Let (Ω, Σ, µ) be a finite measure space, i.e. µ(Ω) < ∞. Then
for p ≥ q ∈ [1, ∞] one has the inclusion
11
1. A Crash Course in Measure Theory
Proof. We only deal with the case p ∈ (1, ∞) (the other cases are easy to show).
It follows from Hölder’s inequality because of p/q ≥ 1 that
Z !1/q Z !1/q Z !1/p Z !(1−q/p)·1/q
q q p
|f | dµ = |f | 1 dµ ≤ |f | dµ 1 dµ
Ω Ω Ω Ω
Z !1/p
p
= µ(Ω)1/q−1/p |f | dµ .
Ω
Note that it is not clear that f ∗ g exists under the above assumptions. This
is indeed the case as the following argument shows. Note that the function
(x, y) 7→ f (y)g(x − y) is measurable as a map R2n → R by the definition of
product σ -algebras and the fact that the product and the composition of
measurable functions is measurable. It follows from Fubini’s theorem that
the function x 7→ (f ∗ g)(x) is measurable and satisfies
Z Z Z Z Z
|f ∗ g| (x) dx ≤ |f (y)||g(x − y)| dy dx = |f (y)| |g(x − y)| dx dy
Rn Z Z Rn R n Rn Rn
kf ∗ gkp ≤ kf k1 kgkp .
Proof. We only deal with the cases p ∈ (1, ∞) as the boundary cases are simple
to prove. We apply Hölder’s inequality
R to the functions |g(x − y)| and 1 for the
measure µ = |f (y)| dy (i.e. µ(A) = A |f (y)| dy) and obtain
Z !1/p Z !1/q
p
|(f ∗ g)(x)| ≤ |g(x − y)| |f (y)| dy |f (y)| dy ,
Rn Rn
12
1.3. Lebesgue Spaces
where 1/p + 1/q = 1. Taking the Lp -norm in the above inequality, we obtain
the desired inequality
Z Z !1/p
p p/q
kf ∗ gkp ≤ |g(x − y)| |f (y)| dy kf k1 dx
Rn Rn
Z Z !1/p
1/q p 1/q 1/p
= kf k1 |f (y)| |g(x − y)| dx dy = kf k1 kf k1 kgkp
Rn Rn
= kf k1 kgkp .
13
The Theory of Self-Adjoint 2
Operators in Hilbert Spaces
2.1 Basic Hilbert Space Theory
By the postulates of quantum mechanics a quantum mechanical system is
described by some complex Hilbert space. Before going any further, we
therefore need some basic results from Hilbert space theory. In this section
we introduce Hilbert spaces and bounded operators between these spaces.
As important examples for the further development, we introduce Fourier
transforms and Sobolev spaces.
We follow the typical physical convention that an inner product on some
complex vector space is linear in the second and anti-linear in the first com-
ponent.
Recall that a sequence (xn )n∈N in a normed space (N , k·k) is called a Cauchy
sequence if for every ε > 0 there exists N ∈ N with kxn − xm k ≤ ε for all n, m ≥ N .
Note that the spaces Cn for n ∈ N are finite-dimensional Hilbert spaces with
respect to the inner product hx|yi = nk=1 xk yk . We now give a first important
P
infinite-dimensional example.
Note that the space L2 (Ω, Σ, µ) is complete by the Riesz–Fischer Theorem 1.3.4.
Further, the finiteness of the scalar product is a consequence of Hölder’s
inequality. As a special case one can take for an open set Ω ⊂ Rn the measure
space (Ω, B(Ω), λ) and obtains the L2 -space L2 (Ω) = L2 (Ω, B(Ω), λ|B(Ω) ).
15
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Definition 2.1.4. A family (ei )i∈I of elements in some Hilbert space H is called
orthogonal if hei |ej i = 0 for all i , j ∈ I. If one additionally has kei k = 1 for all
i ∈ I, one says that (ei )i∈I is orthonormal. If moreover the linear span of (ei )i∈I
is dense in H, the family (ei )i∈I is called an orthonormal basis of H.
16
2.1. Basic Hilbert Space Theory
one sees immediately that FN ◦ f lies in the span of (en )|n|≤N . Hence, the span
of (en )n∈Z is dense in C([0, 1]) and therefore also dense in L2 ([0, 1]) by a variant
of Proposition 1.3.6 (note that L2 ((0, 1)) = L2 ([0, 1])) and we have shown that
(en )n∈Z forms an orthonormal basis of L2 ([0, 1]).
Theorem 2.1.7. Every Hilbert space H has an orthonormal basis. Moreover and
more concretely, if H is infinite dimensional and separable, then there exists an
orthonormal basis (en )n∈N of H.
The proof of the above theorem in the separable case usually uses the
well-known Gram–Schmidt orthonormalization process known from linear
algebra applied to a dense countable subset of H. Orthonormal bases are a
fundamental tool in the study of Hilbert spaces as we see in the following.
Proof. This follows by a direct computation from the relation between the
scalar product and the norm on a Hilbert space. Indeed, we have
kx + yk2 = hx + y|x + yi = hx|xi + hx|yi + hy|xi + hy|yi = kxk2 + kyk2 + hx|yi + hx|yi
= kxk2 + kyk2 + 2 Rehx|yi.
17
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
The uniqueness of the expansion follows directly from the above equation.
For the existence of the expansion we have to show that the partial sums
PN
n=1 hen |xien converge to x in H. By orthogonality we have
2
2
2
XN XN
X N
XN
kxk2 =
x − hen |xien + hen |xien
=
x − hen |xien
+
hen |xien
n=1 n=1
n=1
n=1
2
X N
X N X N
=
x − hen |xien
+ |hen |xi|2 ≥ |hen |xi|2 .
n=1
n=1 n=1
This shows that the sequence (hen |xi)n∈N is square summable and that
∞
X
|hen |xi|2 ≤ kxk2 . (2.1)
n=1
which goes to zero as n → ∞. This shows that the sequence of partial sums
forms a Cauchy sequence in H. By the completeness of H the sequence of
18
2.1. Basic Hilbert Space Theory
for all x ∈ H. This finishes the proof of (a). Now Parseval’s identity in (b) is an
immediate consequence. Indeed, we have for x, y ∈ H
2
X ∞
XN X M
hx|yi =
|hen |xien |
= lim lim hen |xihek |yihen |ek i
n=1 N →∞ M→∞
n=1 k=1
∞
X
= hx|en ihen |yi.
n=1
19
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Corollary 2.1.11 (Plancherel’s identity). Let f ∈ L2 ([0, 1]) and let fˆ(n) for
n ∈ Z denote its n-th Fourier coefficient. Then
Z1 X
|f (x)|2 dx = |fˆ(n)|2 .
0 n∈Z
20
2.1. Basic Hilbert Space Theory
This follows from the fact that (an xn )n∈N ∈ ` 2 (N) whenever (xn )n∈N ∈ ` 2 (N).
Moreover, with the help of the Parseval’s identity one obtains
X ∞ ∞
2 X ∞
X
X∞
2
2 2 2 2
an xn en
=
|an xn | ≤ (sup |an |) |xn | = (sup |an |)
xn en
.
n=1 n=1 n∈N n=1 n∈N n=1
This shows that T is bounded with kT k ≤ supn∈N |an |. Conversely, for all ε > 0
there exists n0 ∈ N such that |an0 | > supn∈N |an | − ε. Let ϕ ∈ [0, 2π) be such that
eiϕ an0 = |an0 |. Then one has keiϕ en0 k = 1 and
Hence, kT k ≥ supn∈N |an | − ε for all ε > 0. Since ε > 0 is arbitrary, the equality
kT k = supn∈N |an | follows. Notice that the same reasoning applies if one
replaces ` 2 by an arbitrary infinite-dimensional separable Hilbert space and
(en )n∈N by an arbitrary orthonormal basis of H.
Notice that the above example in particular shows that in general there
is no x in the unit ball of H such that kT xk = kT k, i.e. T does not attain its
norm. We now turn our attention to dual spaces, a concept fundamental for
the Dirac formulation of quantum mechanics.
21
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
projection onto span{e1 , . . . , eN }. Moreover, one has hPN x|(Id −PN )yi = 0 for all
x, y ∈ H. Indeed, one has
N
*X N
X + N
X * N
X +
hen |xien |y − hek |yiek = hx|en i en |y − hek |yiek
n=1 k=1 n=1 k=1
N
X
= hx|en i(hen |yi − hen |yi) = 0.
n=1
Hence, the kernel and the image of PN are orthogonal subspaces. Such a
projection is called an orthogonal projection.
More generally, let M ⊂ H be a closed subspace of H. Since M is closed, M
is complete with respect to the norm induced by the inherited scalar product
of H. Hence, M is a Hilbert space as well. We assume that M is infinite-
dimensional. The finite dimensional case is simpler and can be treated as
above. Then M has an orthonormal basis (en )n∈N by Theorem 2.1.7. Now
define the linear operator
∞
X
PM : x 7→ hen |xien .
n=1
By estimate (2.1) used in the proof of Theorem 2.1.9 one has kPM xk ≤ kxk.
Moreover, taking again x = e1 we see that PM e1 = e1 . Hence, kPM k = 1. More
generally, one has PM x = x for all x ∈ M as in this case one has x = ∞
P
n=1 hen |xien
2
by Theorem 2.1.9. Furthermore, PM = PM holds and PM is orthogonal. This
can be shown as in the first part of the example. Hence, PM is an orthogonal
projection onto the closed subspace M.
Definition 2.1.15 (Dual space). Let H be a Hilbert space. Then its (topologi-
cal) dual space is defined as
Example 2.1.16. Let H be a Hilbert space. For y ∈ H one defines the func-
tional ϕy (x) = hy|xi on H. It follows from the Cauchy–Schwarz inequality
(Proposition 2.1.3) that for x ∈ H one has |ϕy (x)| ≤ kyk kxk. On the other hand
one has ϕy (y/kyk) = kyk provided y , 0. This shows that ϕy ∈ H0 with norm
kyk (the case y = 0 is obvious).
Remark 2.1.17. The space H0 is called the topological dual space of H because
one requires its elements to be continuous. Sometimes one also considers the
so-called algebraic dual space which consists of all linear functionals H → K.
We will exclusively work with the topological dual space. Hence, no confusion
can arise and we will often drop the term topological.
22
2.1. Basic Hilbert Space Theory
The Riesz representation theorem for Hilbert spaces says that indeed all
elements of H0 are of the form considered in Example 2.1.16.
H → H0 y 7→ ϕy
23
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
denotes by A|ψi the value of ψi under A in agreement with the notation used
in mathematics. One extends the action of A to bras by defining
In fact, if A ∈ B(H) one has |A(hϕ|)(|ψi)| ≤ khϕ|k kAk k|ψik and therefore A(hϕ|) ∈
H∗ for all hϕ| ∈ H∗ .
Now suppose one has given A ∈ B(H) and |ψi ∈ H. Let us determine the
bra hϕ| which corresponds to the ket |ϕi = A|ψi. One has
One obtains hϕ| = hψ|A∗ . Hence, the adjoint formally acts on kets. In particular,
if A is self-adjoint, then A acts in the same way both on kets and bras.
Note that the definition of the action of A on bras is made in a way such
that the value of the scalar product at the right hand side of the above equation
agrees no matter whether A is applied to a ket or to a bra. This justifies the use
of the notation hϕ|A|ψi. Note that if H is finite dimensional and A is identified
with a matrix everything can be computed using matrix multiplications.
One also often uses the so-called outer product of kets and bras. For a ket
|ϕi ∈ H and a bra hψ| ∈ H∗ we define the bounded linear operator
In the finite dimensional setting the operator |ϕihψ| corresponds to the matrix
obtained by multiplying the column vector |ϕi with the row vector hψ|. In
particular, if (en )n∈N is an orthonormal basis of a Hilbert space H, then the
finite rank operator
N
X
|en ihen |
n=1
24
2.1. Basic Hilbert Space Theory
Then one can verify that ψ ∈ C ∞ (R). Now, the function ϕ(x) = ψ(1 + x)ψ(1 − x)
lies in C ∞ (R) and vanishes outside (−1, 1). Hence, ϕ is a non-trivial element
of Cc∞ (R). By taking suitable products of translations and dilations of the
function ϕ just constructed one now easily obtains non-trivial elements of
Cc∞ (Ω).
In the following we use a short-hand notation for higher derivatives. Let
Ω ⊂ Rn be open and α = (α1 , . . . , αn ) ∈ Nn0 be a multi-index. For a suffi-
ciently differentiable function f : Ω → R we write D α f = D α1 D α2 · · · D αn f =
∂|α| f
where |α| = α1 + · · · + αn . Sometimes we will also write Dxα f to
∂α1 x1 ···∂αn xn
,
make clear that we derivative with respect to the x-variables. Moreover, we
use for p ∈ [1, ∞]
p n o
Lloc (Ω) B f : Ω → K measurable : kf 1K kp < ∞ for all compact K ⊂ Ω .
Note that by the Lp -inclusions for finite measure spaces (Proposition 1.3.8)
p q
we have Lloc (Ω) ⊂ Lloc (Ω) for all p ≥ q. In particular, we have the inclusion
p
Lloc (Ω) ⊂ L1loc (Ω) for all p ≥ 1.
The space Cc∞ (Ω) is a so-called space of test functions. Such spaces will
later play an important role in the development of the mathematical theory
of distributions.
In the following example we show that the terminology weak derivative
makes sense.
25
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Example 2.1.22. Consider the function f ∈ L1loc ((−1, 1)) given by f (x) = |x|. It
is well-known that f is not differentiable in the origin. However, g(x) = sign x
is a weak derivative of f . Indeed, for ϕ ∈ Cc∞ ((−1, 1)) one has
Z1 Z1 Z0
0 0
|x| ϕ (x) dx = xϕ (x) dx − xϕ 0 (x) dx
−1 0 −1
Z 1 Z 0
= [xϕ(x)]10 − ϕ(x) dx − [xϕ(x)]0−1 + ϕ(x) dx
0 −1
Z 1
=− sign xϕ(x) dx.
−1
This shows that the weak derivative of f is given by g, i.e. f 0 = g in the weak
sense. Observe that the same argument works if f ∈ L1loc (I) provided one
makes the restriction x0 ∈ I.
Before going any further, we need some approximation results for contin-
uous and Lp functions.
26
2.1. Basic Hilbert Space Theory
Proof. Let ψ ∈ Cc∞ (Rn ) be a non-negative function with kψk1 = 1 and support
inside the unit ball. Now define ψk (x) = k n ψ(kx). Then ψk ≥ 0 and kψk k1 = 1
for all k ∈ N and ψk vanishes outside the ball B(0, 1/k). Now consider the
convolution Z
ϕk (x) = (f ∗ ψk )(x) = f (y)ψk (x − y) dy,
Rn
where we extend f by zero outside Ω. Notice that the convolution exists,
for example as a consequence of Proposition 1.3.10 or by observing that for
fixed x ∈ Rn both sides are integrable after passing to the compact support
of the integrand. It follows from the dominated convergence theorem that
ϕk ∈ C ∞ (Rn ). More precisely, one can verify that for a multi-index α ∈ Nn one
has Z
(D α ϕk )(x) = f (y)(D α ψk )(x − y) dy.
Rn
Moreover, it follows from the formula for the convolution that ϕk (x) vanishes
if |x − y| ≥ 1/k for all y ∈ A. Hence, ϕk vanishes outside A + B(0, 1/k). Since A
has positive distance to ∂Ω, we have ϕk ∈ Cc∞ (Ω) for sufficiently large k.
We now show that (ϕk )k∈N converges to f in Lp . We start with the case
when f additionally is a compactly supported continuous function. Let x ∈ Ω
and ε > 0. Since f is continuous in x, there exists δ > 0 such that |f (x)−f (y)| ≤ ε
for all |x − y| ≤ δ. Now if k > 1/δ
Z Z
|f (x) − ϕk (x)| = f (x)ψk (x − y) dy − f (y)ψk (x − y) dy
Rn R n
Z Z
≤ |f (x) − f (y)|ψk (x − y) dy = |f (x) − f (x − y)|ψk (y) dy
Rn Rn
Z Z
= |f (x) − f (x − y)|ψk (y) dy ≤ ε ψk (y) dy = ε.
B(0,1/k) B(0,1/k)
27
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Since the functions ϕk are supported in A + B(0, 1) and are uniformly bounded
by kf k∞ (which is finite because f is continuous and has compact support), it
follows from the dominated convergence theorem that ϕk → f in Lp .
We now consider the case of arbitrary f ∈ Lp (Ω). Consider for k ∈ N the
linear bounded operator
Tk : Lp (Ω) → Lp (Rn )
f 7→ f ∗ ψk
Notice that a priori a function f could have several different weak deriva-
tives. We now show that this is not the case. The following lemma is often
called the du Bois-Reymond lemma. In calculus of variations you have probably
encountered variants of this lemma which are usually called the fundamental
lemma of calculus of variations. There it is used to deduce the Euler–Lagrange
equations from the variational principle. Although intuitively clear, a rigorous
proof needs some effort because of measure theoretic difficulties.
Lemma 2.1.25 (du Bois-Reymond). Let Ω ⊂ Rn be open and f ∈ L1loc (Ω) with
Z
fϕ=0 for all ϕ ∈ Cc∞ (Ω).
Ω
Proof. First observe that it is sufficient to consider the case when f is a real
function. Indeed, the assumption implies
Z Z Z
fϕ= Re f ϕ + i Im f ϕ = 0
Ω Ω Ω
for all real ϕ ∈ Cc (Ω). Since a complex number vanishes if and only if both
the real and imaginary part vanish, both summands in the above formula
must vanish. Hence, the complex case follows from the real case applied to
both Re f and Im f .
For n ∈ N let Ωn = {x ∈ Ω∩B(0, n) : dist(x, ∂(Ω∩B(0, n)) > 1/n}. Then Ωn is
open and bounded. Suppose we can show that f = 0 almost everywhere on Ωn .
Then it follows that f = 0 almost everywhere on Ω because of ∪n∈N Ωn = Ω
and the fact that the countable union of null sets is a null set.
Now assume that f = 0 does not hold almost everywhere on Ωn . This
means that |{x ∈ Ωn : |f (x)| > 0}| > 0. We may assume without loss of generality
28
2.1. Basic Hilbert Space Theory
that |{x ∈ Ωn : f (x) > 0}| > 0 (replace f by −f if necessary). It now follows that
there exists an ε > 0 and a measurable subset A ⊂ Ωn of positive measure with
f (x) ≥ ε for all x ∈ A.
Let B be an arbitary measurable subset of Ωn . Since Ωn has positive dis-
tance to the boundary of Ω, by Proposition 2.1.24 there exists a sequence
(ϕk )k∈N ⊂ Cc∞ (Ωn+1 ) with 0 ≤ ϕk ≤ 1 and ϕk → 1B in L1 (Ω). By Proposi-
tion 1.3.5 we can additionally assume after passing to a subsequence that
ϕk (x) → 1B (x) almost everywhere on Ω. Since Ωn has finite measure, it fol-
lows from the dominated convergence theorem because of |ϕk f | ≤ |f | 1Ωn+1
that Z Z
f 1B = lim f ϕk = 0.
Ω k→∞ Ω
Taking B = A, we however have by the considerations in the previous para-
graph Z Z
f 1A ≥ ε 1A = ε |A| > 0,
Ω Ω
which is a contradiction. Hence, we must have f = 0 almost everywhere on
Ωn .
Now suppose that f ∈ L1loc (Ω) has a weak α-th derivative. By the above
corollary a weak derivative is uniquely determined in L1loc (Ω). It makes
therefore sense to speak of the weak α-th derivative which we will denote
by D α f . Recall that we have seen in Example 2.1.21 that if f is classically
continuously differentiable, the weak derivative coincides with the classical
derivative. Thus there is no conflict in notation.
We now can finally define Sobolev spaces.
29
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
W k,p (Ω) B {f ∈ Lp (Ω) : f is weakly α-diff. and D α f ∈ Lp (Ω) for all |α| ≤ n}.
which endows W k,2 (Ω) with the structure of a Hilbert space. We will also use the
abbreviation H k (Ω) = W k,2 (Ω).
Proof. All assertions except for the completeness are obvious. For the com-
pleteness suppose that (fn )n∈N is a Cauchy sequence in W k,p (Ω). In particular,
(D α fn )n∈N is a Cauchy sequence in Lp (Ω) for all |α| ≤ k. Now, it follows from
the completeness of Lp (Ω) that for all |α| ≤ k there exists fα ∈ Lp (Ω) with
D α fn → fα .
We now show that f = f(0,...,0) is weakly α-differentiable for all |α| ≤ k. For
this observe that for all ϕ ∈ Cc∞ (Ω) one has
Z Z Z Z
α α |α| α |α|
f D ϕ = lim fn D ϕ = lim (−1) D fn ϕ = (−1) fα ϕ.
Ω n→∞ Ω n→∞ Ω Ω
R
Here we have used twice the fact that for ψ ∈ Cc∞ (Ω) the functional f 7→ Ω f ψ
is continuous by Hölder’s inequality. Observe that the above calculation
shows that D α f = fα . From this it is now clear that f ∈ W k,p (Ω) and fn → f in
W k,p (Ω). Hence, W k,p (Ω) is complete.
Note that we have seen in Example 2.1.22 that there exist Sobolev functions
which are not classically differentiable. However, one has the following
denseness result.
Theorem 2.1.29. Let n ∈ N, p ∈ [1, ∞) and k ∈ N. Then the space Cc∞ (Rn ) is
dense in W k,p (Rn ).
30
2.1. Basic Hilbert Space Theory
Proof. Let f ∈ W k,p (Rn ). By Proposition 2.1.24 the sequence in Cc∞ (Rn ) given
by ϕk = f ∗ ψk , where ψk (x) = k n ψ(kx) for some Cc∞ (Rn ) with kψk1 = 1, con-
verges to f in Lp . Moreover, we have seen that for α ∈ Nn
Z
(D α ϕk )(x) = f (y)(Dxα ψk )(x − y) dy.
Rn
Observe that (Dxα ψk )(x − y) = (−1)|α| (Dyα ψk )(x − y) lies in Cc∞ (Rn ). Hence, for
|α| ≤ k we have by the definition of the weak derivative
Z Z
(D α ϕk )(x) = (−1)−α f (y)(Dyα ψk )(x − y) dy = (D α f )(y)ψk (x − y) dy
Rn Rn
α
= D f ∗ ψk
31
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
w ∈ Cc∞ (Rn ). Note that if u is a classical solution such that f as well as u and
all of its partial derivatives of first order lie in L2 (Rn ), then (WS) holds for all
w ∈ H 1 (Rn ) by the density of Cc (Ω) in H 1 (Rn ). More generally, the validity of
(WS) extends from all w ∈ Cc∞ (Rn ) to all w ∈ H 1 (Rn ) provided u ∈ H 1 (Rn ) and
f ∈ L2 (Rn ).
The advantage of the concept of weak solutions lies in the fact that one
often can establish the existence and uniqueness of weak solutions with
functional analytic methods. In our case consider the functional
ϕ : H 1 (Rn ) → R
Z
w 7→ f w dx.
Rn
for all w ∈ H 1 (Rn ). This shows that ϕ ∈ H 1 (Rn )∗ and it now follows from
the Riesz representation theorem (Theorem 2.1.18) that there exists a unique
u ∈ H 1 (Rn ) such that
Z Z Z
hu, wiH 1 (Rn ) = ∇u∇w dx + uw dx = f w dx = ϕ(w)
Rn Rn Rn
for all w ∈ H 1 (Rn ). Hence, we have shown that there exists a unique weak
solution u of the equation for all f ∈ L2 (Rn ). We will later see that this
solution already satisfies u ∈ C ∞ (Rn ) if the inhomogeneity additionally has
the regularity f ∈ C ∞ (Rn ) and therefore is a classical solution of −∆u + u = f .
32
2.1. Basic Hilbert Space Theory
Observe that C0 (Ω) becomes a Banach space when endowed with the norm
kf k∞ B supx∈Ω |f (x)|. This follows from the fact that the uniform limit of
continuous functions is continuous and respects the vanishing condition. The
Fourier transform has the following elementary but useful mapping property.
The above lemma can be explicitly verified for indicators of finite intervals
and hence extends to simple functions by linearity. The general case then
follows from a density argument and the fact that C0 (Rn ) is a closed subspace
of L∞ (Rn ). Note that for a function f ∈ L2 (Rn ) the above Fourier integral may
not converge. Nevertheless it is possible to extend the Fourier transform to
L2 (Rn ).
4
1X k
hx|yi = i kx + i k yk2
4
k=1
(a) Let f ∈ L1 (Rn ) such that F f ∈ L1 (Rn ) as well. Then the Fourier inversion
formula Z
1
f (x) = (F f )(y)eix·y dy
(2π)n/2 Rn
holds.
33
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
More generally, one has f ∈ H k (Rn ) if and only if xα F f lies in L2 (Rn ) for all
|α| ≤ k. More precisely, for α ∈ Nn the weak α-th partial derivative exists and lies
in L2 (Rn ) if and only if xα F f ∈ L2 (Rn ). In this case
Proof. First let f ∈ Cc∞ (Rn ). Using integration by parts we obtain for j = 1, . . . , n
and x ∈ Rn the identity
!! Z Z
∂f 1 ∂f −ixy 1
F (x) = (y)e dy = ixj f (y)e−ixy dy
∂yj (2π)n/2 Rn ∂yj (2π)n/2 Rn
= ixj (F f )(x).
∂f
Hence, F ( ∂y ) = ixj F f . Since the space of test functions Cc∞ (Rn ) is dense
j
exists in L2 (Rn ). It follows from Proposition 1.3.5 that limn→∞ ixj F fn agrees
with ixj F f almost everywhere. Hence, ixj F f ∈ L2 (Rn ) with F (Dj f ) = ixj F f .
34
2.1. Basic Hilbert Space Theory
=− F f F (Dj ϕ) = − f Dj ϕ = − f Dj ϕ.
Rn Rn Rn
Hence, by definition f ∈ H 1 (Rn ). The general case and the norm identity
follow on exactly the same lines of proof.
Theorem 2.1.36 (Sobolev embedding). Let k, n ∈ N with k > n2 . Then one has
the inclusion H k (Rn ) ⊂ C0 (Rn ). Hence, H k (Rn ) ⊂ C m (Rn ) ∩ C0 (Rn ) if k − m > n2 .
In particular ∩k∈N H k (Rn ) ⊂ C ∞ (Rn ).
Proof. Let k, n ∈ N be such that 2k > n. It follows from Proposition 2.1.35 that
|x|α F f ∈ L2 (Rn ) for all |α| ≤ k. It then follows that for some constants cα
Z X Z
2
2 k
(1 + |x| ) |(F f )(x)| dx = cα (xα )2 |(F f )(x)|2 dx < ∞.
Rn |α|≤k Rn
This shows (1 + |x|2 )k/2 F f ∈ L2 (Rn ). Now we obtain with the Cauchy–
Schwarz inequality that
By the above calculation the first factor factor is finite, whereas the second
factor is finite because of the assumption k > n2 (note that (1 + |x|2 )−α is inte-
grable if and only if 2α > n). Hence, F f ∈ L1 (Rn ). It now follows from the
Riemann–Lebesuge lemma (Lemma 2.1.32) that f = F −1 F f ∈ C0 (Rn ).
For the proof of the higher inclusions observe that if k − 1 > n2 we have
Dj f ∈ H k−1 (Rn ) for all j = 1, . . . , n. Applying the just shown result to the
partial derivatives, we obtain Dj f ∈ C0 (Rn ). This shows f ∈ C 1 (Rn ). The
general case now follows inductively.
kf kC m (Ω ≤ kf kW k,p (Ω) .
35
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
In fact, the existence of such a constant follows abstractly from the em-
bedding result by the closed graph theorem (Theorem 2.2.9) that will be
introduced later in the lecture. The Fourier transform has a lot of important
applications as it gives a very easy description of very important operators in
analysis. We illustrate this fact with an easy example.
for all test functions ϕ ∈ Cc∞ (Rn ). In analogy to the notion of weak derivatives,
we say a function w ∈ L1loc (Rn ) is the (unique) weak Laplacian of some function
u ∈ L1loc (Rn ) provided the integration by parts formula
Z Z
u∆ϕ = wϕ
Rn Rn
holds for all test functions ϕ ∈ Cc∞ (Rn ). Using this terminology for the weak
solutions of −∆u + u = f , we see that because of
Z Z Z
− u∆ϕ = ∇u∇ϕ = (f − u)ϕ
Rn Rn Rn
for all ϕ ∈ Cc∞ (Rn ) that the weak Laplacian of u lies in L2 (Rn ) and satisfies
−∆u = f − u. Since ∆u ∈ L2 (Rn ), it follows from Proposition 2.1.35 that |x|2 F f
and therefore also (1 + |x|2 )F f is square integrable. We now want to show
that all mixed second partial derivatives exist and are square integrable. By
Proposition 2.1.35 this is equivalent to xi xj F f ∈ L2 (Rn ) for all i, j = 1, . . . , n.
Fix such i, j. Then
2
xi xj
Z 2 Z 2
xi xj (F u)(x) dx = 2
2 (1 + |x| )(F u)(x) dx
Rn Rn 1 + |x|
2 Z
xi xj
≤ sup 2
(1 + |x|2 )2 |(F u)(x)|2 dx ≤ k∆uk2L2 (Rn ) .
x∈Rn 1 + |x| Rn
A similar estimates holds for the first derivatives. Hence, u ∈ H 2 (Rn ) whenever
∆u ∈ L2 (Rn ). In particular, the weak solution u ∈ H 1 (Rn ) of −∆u + u = f
automatically has the higher regularity u ∈ H 2 (Rn ). Now suppose that the
inhomogeneity even satisfies f ∈ H 1 (Rn ). Then ∆u = u − f ∈ H 1 (Rn ). Using
a variant of the above arguments we see that u ∈ H 3 (Rn ). Iterating this
36
2.2. Symmetric and Self-Adjoint Operators
argument we see that f ∈ H k (Rn ) implies the higher regularity u ∈ H k+2 (Rn ).
This is the so-called elliptic regularity of the Laplace operator. Note that this in
particular implies by the Sobolev embedding Theorem 2.1.36 that u ∈ C ∞ (Rn )
whenever f ∈ C ∞ (Rn ) and additionally has suitable decay of all derivatives
such that f lies in Sobolev spaces of arbitrary high order.
d
x̂ : f 7→ [x 7→ x · f (x)] p̂ : f 7→ [x 7→ i f (x)].
dx
Clearly, both x̂ and p̂ are linear. However, they do not define bounded operators
on L2 (R). There are two (closely related) obstructions:
1. x̂(f ) does not lie in L2 (R) for all f ∈ L2 (R) (for example one can take
f (x) = 1x 1[1,∞) ). In the same spirit not every function in L2 (R) has a
(weak) derivative in L2 (Rn ).
2. Both operators are not bounded when restricted to their maximal do-
mains of definition. For example, one has
sup kx̂(f )k = ∞.
kf k2 ≤1:xf ∈L2 (R)
In particular, we say that two unbounded operators (A, D(A)) and (B, D(B))
agree and one writes A = B if and only if D(A) = D(B) and Ax = Bx for
all x ∈ D(A) = D(B). If one has D(A) ⊂ D(B) and Bx = Ax for all x ∈ D(A),
we say that B is an extension of A. Moreover, we define the sum of two
unbounded operators (A, D(A)) and (B, D(B)) on the same Hilbert space H
in the natural way: (D(A + B), A + B) is given by D(A + B) = D(A) ∩ D(B) and
(A + B)x B Ax + Bx. In particular, the sum of an unbounded operator (A, D(A))
37
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Definition 2.2.4 (Spectrum and resolvent set of an operator). Let (A, D(A))
be an (unbounded) operator on some Hilbert space H. We call the set
ρ(A) B {λ ∈ K : λ Id −A is invertible}
the resolvent set of A. Its complement σ (A) B K \ ρ(A) is called the (mathemat-
ical) spectrum of A.
38
2.2. Symmetric and Self-Adjoint Operators
(T f )(x) = xf (x).
so f < L2 ([0, 1]). With similiar arguments one can further show that [0, 1] ⊂
σ (A). Note that on a (until now) formal level the Dirac measure δ0 at zero is
an eigenvalue of T because of
T δ0 = xδ0 = 0.
HereRxδ0 is the Dirac measure with the density x, i.e. the Borel measure
A 7→ A x dδ0 . But please be aware that δ0 is not an element of L2 ([0, 1]) as it is
not even a function. At the end of the lecture we will give mathematical sense
to such expressions by introducing the theory of distributions. Moreover, note
that the same discussion applies to the unbounded position operator on L2 (R),
where one even obtains R ⊂ σ (A).
We will almost exclusively work with the class of closed operators on some
Hilbert space.
Observe that if (A, D(A)) is an injective closed operator, then the graph of
A−1 is given by G(A−1 ) = {(y, x) : (x, y) ∈ G(A)}. This shows that an injective
operator A is closed if and only if A−1 is closed. In particular, if ρ(A) is non-
empty, then λ − A is invertible for some λ ∈ C. Hence, (λ − A)−1 is bounded
39
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
and a fortiori closed. Now, it follows that λ − A and therefore also A are closed.
This shows that ρ(A) , ∅ implies that A is closed. Hence, a reasonable spectral
theory is only possible for closed operators.
As a consequence we only encounter closed operators in our study of
the mathematics behind quantum mechanics. Indeed, by the postulates of
quantum mechanics a physical quantum mechanical system is modeled by
a self-adjoint operator on some complex Hilbert space H. The mathematical
spectrum σ (H) of H corresponds by the postulates of quantum mechanics to
the possible outcomes of a physical measurement of the system. Naturally,
one requires σ (H) ⊂ R (reality condition). A fortiori one has ρ(A) , ∅. In other
words, there exists a λ ∈ C such that (λ − A)−1 : H → H is a bounded operator.
By the above reasoning this implies that A must be closed.
Remark 2.2.7. One can easily verify that the definition of a closed operator
is equivalent to the following condition: for (xn )n∈N ⊂ D(A) with xn → x and
Axn → y in H one has x ∈ D(A) and Ax = y.
where the derivative of f is understood in the weak sense. From this one sees
that (fn , Afn ) ∈ G(A) with (fn , Afn ) → (f , f 0 ) in H×H. Hence, G(A) is not closed
in H × H and (A, D(A)) does not define a closed operator.
d
However, if we choose B = dx with D(B) = H 1 ((0, 1)), one can immediately
verify that B is a closed operator on H. Moreover, it follows from the above
calculations that G(A) ⊂ G(B). From this one sees that A is closable and that B
is an extension of A.
40
2.2. Symmetric and Self-Adjoint Operators
In particular notice that in the above example one only obtains a closed
operator if the domain consists of Sobolev and not classically differentiable
functions. This is the prototypical behaviour for all differential operators.
One has the following fundamental theorem on closed operators.
Theorem 2.2.9 (Closed Graph Theorem). Let (A, D(A)) be a closed operator on
some Hilbert space H with D(A) = H. Then A is bounded, i.e. A ∈ B(H).
Let us briefly discuss the consequences of the closed graph theorem for the
mathematical description of quantum mechanics. A typical quantum mechan-
ical system is described by a closed unbounded operator H (the Hamiltonian)
on some Hilbert space H. The closed graph theorem implies that one then
automatically has D(H) ( H, i.e. one has to restrict the domain of the operator.
41
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Z ∞
= if 0 (x)g(x) dx = hAf |gi.
0
This shows that A is symmetric. However, it follows along the same line of
arguments as in Example 2.2.8 that A is not closed. Hence, by the comments
after Definition 2.2.6 one has σ (A) = C.
Before going further, we prove a very useful lemma for symmetric opera-
tors.
Lemma 2.2.12. Let (A, D(A)) be a symmetric operator on some Hilbert space H.
Then for all λ ∈ C we have
The above lemma has the following very useful consequences which sim-
plify the study of symmetric operators substantially.
Corollary 2.2.13. Let (A, (D(A)) be a symmetric operator on some Hilbert space
H. Then the following are equivalent for λ ∈ C \ R.
(i) λ ∈ ρ(A);
(ii) λ − A is surjective.
Proof. The first condition implies the first by definition. For the converse
observe that λ − A is injective by Lemma 2.2.12 or the fact that a symmetric
operator does not have any eigenvalues. Hence, λ − A is bijective. Now, it
follows from the estimate in Lemma 2.2.12 that for y ∈ H with (λ − A)x = y we
have
k(λ − A)−1 yk = kxk ≤ |Im λ|−1 k(λ − A)xk = |Im λ|−1 kyk.
42
2.2. Symmetric and Self-Adjoint Operators
Corollary 2.2.14. Let (A, D(A)) be a symmetric operator on some Hilbert space
H. If λ − A is surjective for some λ ∈ C \ R, then D(A) is dense in H.
Proof. Assume that this is not the case. Then the closure D(A) of H is a proper
subspace of H. Hence, there exists y ∈ H \ D(A). Let P be the orthogonal
projection onto D(A). Replacing y by y − P y if necessary, we may assume
that y is orthogonal to all elements of D(A). By assumption, there exists
0 , x ∈ D(A) with (λ − A)x = y. Now, on the one hand
Note that by Remark 2.1.37 one has the Sobolev embedding H 1 ((0, 1)) ⊂
C([0, 1]). More precisely, this means that every function f ∈ H 1 ((0, 1)) has an
(automatically unique) representative in C([0, 1]). Note that in particular it
then makes sense to talk about point evaluations of f , e.g. f (0) or f (1). Please
remember that point evaluations are not well-defined for general functions
in L2 ([0, 1]). The above inclusion into the continuous functions allows us to
define the following subspace of H 1 ((0, 1)).
Definition 2.2.15. Let a < b ∈ R. Then we define the Sobolev space with
vanishing boundary values as
In an analogous way one also defines the spaces H01 ((−∞, b]).
Remark 2.2.16. Using the methods of the proof of Theorem 2.1.36 and a trun-
cation argument one can show that Cc ((0, ∞)) is a dense subset of H01 ([0, ∞)).
Analogously, Cc ((a, b)) is a dense subset of H01 ((a, b)).
In Example 2.2.11, one may argue that the essential obstruction is the fact
that A is not closed. Indeed, this fact was used to show σ (A) = C. However,
there are even more obstructions.
43
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Example 2.2.17 (Example 2.2.11 continued). Now consider the same map-
d
ping A = i dx , but with the different domain D(A) = H01 ([0, ∞)). Now the
derivative has to be understood in the weak sense. Then A is a closed operator.
This can be checked as in the previous examples: for the additional prob-
lem of the boundary values please observe that fn → f in H 1 ((0, ∞)) implies
fn (x) → f (x) for all x ∈ [0, ∞) by the estimate for the Sobolev embedding
theorem (see end of Remark 2.1.37). It is not clear that A is still symmetric
on the bigger domain which we use now. Until now, we only know from
Example 2.2.11 that
hf |Agi = hAf |gi
for all f , g ∈ Cc∞ ((0, ∞)). Now let f , g ∈ H01 ([0, ∞)). Since this space is dense
by the remarks made before this example, there exists sequences (fn )n∈N and
(gn )n∈N in Cc∞ ([0, ∞)) such that fn → f and gn → g in H01 ([0, ∞)). A fortiori,
we have fn → f and gn → g in L2 ([0, ∞)). Moreover, for h ∈ H01 ([0, ∞)) we have
This shows that A is continuous as a mapping from D(A) = H01 ([0, ∞)) to
L2 ([0, ∞)). In particular fn → f in H01 ([0, ∞)) implies Afn → Af in L2 ([0, ∞)).
Using the continuity of the scalar product we obtain
Hence, (A, D(A)) is still symmetric. Note that loosely spoken the symmetry
gets increasingly difficult to achieve for bigger domains. For example, if we
would use D(A) = H 1 ([0, ∞)) instead, in the integration by parts argument the
term f (0)g(0) would not vanish in general. Hence, A is not symmetric with
this even bigger domain.
We now deal with the spectrum of A. Let us first consider λ ∈ C with
Im λ < 0. Note that by Corollary 2.2.13 the essential question is whether λ − A
is surjective. For this let g ∈ L2 ([0, ∞)). We have to find a (necessarily unique)
f ∈ H01 ([0, ∞)) with
λf − Af = λf − if 0 = g.
Using the well-known variation of parameters method we can write down the
very reasonable candidate (ignoring problems such as weak differentiation
for a moment) Z t
f (t) = i g(s)e−iλ(t−s) ds.
0
We now check that f is square integrable. For this observe that by the Cauchy–
Schwarz inequality
Zt !2 Z t Zt
2 Im λ(t−s) 2 Im λ(t−s)
|f (t)| ≤ |g(s)| e ds ≤ |g(s)| e ds eIm λ(t−s) ds
0 0 0
44
2.2. Symmetric and Self-Adjoint Operators
Z t
1
≤ |g(s)|2 eIm λ(t−s) ds.
|Im λ| 0
Hence, f ∈ L2 ([0, ∞)) with kf k2 ≤ |Im λ|−1 kgk2 . Now a calculation similar to
that of Example 2.1.23 shows that f is weakly differentiable with if 0 = λf − g.
It follows that f 0 ∈ L2 ([0, ∞)) and therefore f ∈ H 1 ([0, ∞)). In particular, f is
continuous (this can also be easily verified directly by using the dominated
convergence theorem) with f (0) = 0. Altogether this shows that f ∈ H01 ([0, ∞))
with (λ − A)f = g. Corollary 2.2.13 shows that λ ∈ ρ(A) whenever Im λ < 0.
Hence, the upper half-plane is contained in ρ(A).
However, the situation is different for the upper half-plane. For example,
take λ = i and g(t) = e−t ∈ L2 ([0, ∞)). In this case the resolvent equation is the
ordinary differential equation −if (t) − if 0 (t) = e−t with the initial condition
f (0) = 0 whose unique solution (note that the right hand side lies in Sobolev
spaces of arbitrary order, by bootstrapping this shows that a solution f ∈
H 1 ([0, ∞)) automatically satisfies f ∈ C ∞ ((0, ∞)) and therefore is a classical
solution of the ODE) is given by
Zt Zt
i
f (t) = i e−s et−s ds = iet e−2s ds = et (1 − e−2t )
0 0 2
45
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
d
Indeed, we have shown that the momentum operator p̂ = −i dx on the half-line
[0, ∞) with zero boundary condition is not self-adjoint and therefore is not a
physical reasonable observable. Even worse, we will later show that the oper-
d
ator −i dx defined on Cc∞ ((0, ∞)) has no self-adjoint extensions. Clearly, if ψ is
a quantum mechanical wave function which is localized in a compact subset
of (0, ∞), the time evolution of the system should be given for sufficiently
small times by the time evolution of a free particle, i.e. by translation as we
have seen in the physics part of the lecture. Hence, on Cc∞ ((0, ∞)) the single
d
choice we have is to set p̂ = −i dx . But since this operator does not have any
self-adjoint extensions, by the postulates of quantum mechanics, there does
not exist a well-defined momentum observable. Clearly, this odd behaviour
should arise from the infinite high well at zero.
Since the well is infinitely high, we would expect that a wave function
ψ is totally reflected at zero (or at least with some phase shift), i.e. the time
evolution gets an immediate phase shift of π. You probably all know this kind
of reasoning from classical mechanics where this argument works perfectly
fine. But now for simplicity think of a wave function of the form ψ(x) = 1[0,1] .
Then one has kψk = 1. Let us for simplicity assume that the time evolution
operators U (t) of the momentum operator are given by translation of one per
time unit. Assuming total reflection, the state has evolved after a time span of
1/2 to the state U (1/2)ψ which satisfies
Hence, U (t) does not preserve the norm of the state and therefore violates a
fundamental principle of quantum mechanics. Again physical and mathemat-
ical reasoning fits perfectly together!
The occurrence of half-planes as the spectrum of symmetric operators is
no coincidence as the next result shows. Before we introduce the resolvent of
an operator and need some technical tools.
Definition 2.2.19 (Resolvent). Let (A, D(A)) be an unbounded operator on
some Hilbert space H. Then the mapping
46
2.2. Symmetric and Self-Adjoint Operators
The next elementary result plays a fundamental role in the study of opera-
tors on Banach or Hilbert spaces.
which exists because of kT k < 1. The first assertion now follows by taking
limits. For the second assertion write
S = T + S − T = T (Id +T −1 (S − T )).
kT −1 (S − T )k ≤ kT −1 k kT − Sk < 1.
Note that it follows from the above lemma that the set of invertible opera-
tors is open in B(X) for every Banach space X. Moreover, one obtains directly
the previously stated fact that the resolvent set ρ(A) of an unbounded operator
(A, D(A)) is open (and as a consequence that the spectrum σ (A) is closed). We
are now ready to describe the spectrum of a symmetric operator.
Proposition 2.2.21. Let (A, D(A)) be a symmetric operator on some Hilbert space
H. If λ Id −A is surjective for some λ ∈ C with Im λ > 0 (resp. Im λ < 0), then the
whole upper (resp. lower) half-plane is contained in the resolvent set ρ(A) of A.
47
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Hence, it follows from the Neumann series (Lemma 2.2.20) that the right
hand side and therefore µ − A is surjective and consequently µ ∈ ρ(A) by
Corollary 2.2.13 provided
1
|µ − λ| kR(λ, A)k < 1 ⇔ kR(λ, A)k < .
|µ − λ|
This is satisfied if
1 1
< ⇔ |µ − λ| < |Im λ| .
|Im λ| |µ − λ|
Hence, for every λ ∈ ρ(A) the ball with center λ and radius |Im λ| is contained
in ρ(A) as well. By taking bigger and bigger disks we see that the whole upper
half plane is contained in the resolvent set ρ(A).
Corollary 2.2.22. Let (A, D(A)) be a symmetric operator on some Hilbert space
H. Then the spectrum σ (A) of A is given by one of the following four possibilities:
a closed subset of the real line, the upper or lower closed half-plane or the entire
complex plane.
Proof. By Lemma 2.2.21 the resolvent set ρ(A) contains a whole open half-
plane as soon as a single point of this half-plane is contained in ρ(A). Passing
to complements this means that either {λ ∈ C : Re λ > 0} or no point of this
half-plane are contained in the spectrum σ (A). The same of course applies
the lower half-plane. Since the spectrum σ (A) is closed, only the above listed
cases remain possible.
In fact, one can construct for each of the above listed closed sets a closed (!)
densely defined symmetric operator on some Hilbert space whose spectrum is
exactly that set. Hence, the above statement is the best possible in the general
case of a symmetric operator. In order to guarantee that the spectrum of an
unbounded operator is real we therefore need a stronger condition than mere
symmetry. It is now finally time to introduce the adjoint of a densely defined
operator on some Hilbert space.
Definition 2.2.23. Let (A, D(A)) be a densely defined operator on some Hilbert
space H. The adjoint A∗ of A is the operator defined by
48
2.2. Symmetric and Self-Adjoint Operators
ϕ : D(A) → H
z 7→ hx|Ayi
is continuous, i.e. there exists a constant C ≥ 0 such that |ϕ(z)| ≤ C kzkH for all
z ∈ D(A). Since D(A) is dense in H, there exists a unique extension ϕ̃ ∈ H∗ of
ϕ to all of H. In fact by the definition, one has ϕ̃(z) = hg|zi for all z ∈ H. Note
that by the Riesz representation theorem (Theorem 2.1.18) each continuous
functional on H can be uniquely determined by such an element in H, in our
case by the element y. Hence, an equivalent definition of the domain is
Note that by the above definition the element y is uniquely determined. This
shows that A∗ is well-defined. Moreover, it follows from a direct computation
that A∗ is indeed linear. Further, the adjoint is always closed.
Lemma 2.2.24. Let (A, D(A)) be a densely defined operator on some Hilbert space
H. Then the adjoint operator (A∗ , D(A∗ )) is closed.
Proof. Let (xn )n∈N ⊂ D(A∗ ) with xn → x and A∗ xn → y ∈ H. Then one has for
all z ∈ D(A)
hx|Azi = lim hxn |Azi = lim hA∗ xn |zi = hy|zi.
n→∞ n→∞
Hence, x ∈ D(A∗ ) and A∗ x = y. This shows that A∗ is a closed operator.
Example 2.2.25. Recall that the operator from Example 2.2.11 was given by
d
A = i dx with domain D(A) = H01 ([0, ∞)). Assume that f ∈ D(A∗ ). Then we
have for all g ∈ D(A)
Z∞ Z∞
∗
∗
A f g = hA f |gi = hf |Agi = f (x)ig 0 (x) dx
0 0
In particular, if we only consider real ϕ = g ∈ Cc∞ ([0, ∞)) and use the anti-
symmetry of the scalar product, we obtain
Z∞ Z∞
∗
− iA f ϕ = − f (x)ϕ 0 (x) dx.
0 0
49
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
d
H 1 ([0, ∞)) and that A∗ is a restriction of the differential operator i dx with
1 ∗ 1
domain H ([0, ∞)). We now show that D(A ) = H ([0, ∞)). For this let
f ∈ H 1 ([0, ∞)). Then one has for all g ∈ D(A) = H01 ([0, ∞)) using integration by
parts that
Z ∞ h i∞ Z ∞
0
hf |Agi = f (x)ig (x) dx = if (x)g(x) − f 0 (x)ig(x) dx
0
Z0∞ 0
Note that equality relying upon integration by parts holds for all test functions
g ∈ Cc∞ ((0, ∞)) and extends to all g ∈ H01 [0, ∞) by continuity because the
test functions Cc∞ ([0, ∞)) are dense in H01 ([0, ∞)) as stated in Remark 2.2.16.
The above identity implies that H 1 ([0, ∞)) ⊂ D(A∗ ) and A∗ f = if 0 for all
d
f ∈ H 1 ([0, ∞)). Altogether, we have shown that A∗ = i dx with domain D(A∗ ) =
H 1 ((0, ∞)). Hence, we do not have A = A∗ . However, A∗ is an extension of A.
Note that (A∗ , D(A∗ )) is not symmetric although (A, D(A)) is symmetric.
Indeed, let f (x) = g(x) = e−x/2 ∈ H 1 ([0, ∞)). Then by integration by parts we
see that the symmetry is violated as
Z ∞ Z ∞
∗ 0 −x
hf |A gi = hf |ig i = −i e dx = i + i e−x dx = i + hif 0 |gi = i + hA∗ f |gi.
0 0
Lemma 2.2.26. Let (A, D(A)) be a densely defined symmetric operator on some
Hilbert space. Then (A∗ , D(A∗ )) is a closed extension of A. In particular, (A, D(A))
is closable.
Proof. We have seen in Lemma 2.2.24 that the adjoint is a closed operator. It
remains to show that A∗ extends A. Let x ∈ D(A). Then one has for all y ∈ D(A)
by the symmetry of A
hx|Ayi = hAx|yi.
50
2.2. Symmetric and Self-Adjoint Operators
Using this definition, we have shown in Example 2.2.25 that the operator
d
A = i dx with domain D(A) = H 1 ([0, ∞)) is not self-adjoint as the adjoint is a
true extension of A.
We now give some basic positive examples of self-adjoint operators.
Example 2.2.28 (Multiplication Operators). Let (Ω, Σ, µ) be a measure space
and m : Ω → R a measurable function. One defines the multiplication opera-
tor Mm on L2 (Ω) = L2 (Ω, Σ, µ) as
D(Mm ) = {f ∈ L2 (Ω) : m · f ∈ L2 (Ω)}
Mm f = m · f .
First observe that Mm is densely defined. For this consider the set An = {x ∈ Ω :
|m(x)| ≤ n} for n ∈ N. Since m is bounded on An , one has f 1An ∈ D(Mm ) for all
f ∈ L2 (Ω) and all n ∈ N. Moreover, it follows from the dominated convergence
theorem that f 1An → f in L2 (Ω) for all f ∈ L2 (Ω). Hence, D(Mm ) = L2 (Ω).
Note that since D(Mm ) is dense, we can now speak of the adjoint.
Now let us compute the adjoint of Mm . First assume that f , g ∈ D(Mm ).
Then one has
Z Z
hf |Mm gi = f mg dµ = mf g dµ = hMm f |gi.
Ω Ω
This shows that Mm is symmetric and (therefore automatically) Mm is a restric-
tion of (Mm )∗ . It remains to show that the converse inclusion Mm ∗ = (M )∗ ⊆
m
Mm holds as well. Suppose that f ∈ D(Mm ∗ ). Then one has for all g ∈ D(M )
m
by definition of the adjoint
Z Z Z
∗ ∗ ∗
Mm f g dµ = hMm f |gi = hf |Mm gi = f mg dµ ⇔ Mm f − mf g dµ = 0.
Ω Ω Ω
and obtain Z
|Mm ∗
f − mf |2 dµ = 0.
Ω
This shows that Mm= mf . In particular, mf ∈ L2 (Ω) and therefore f ∈
∗f
Note that if we choose (R, B(R), λ) as measure space and m(x) = x as multi-
plier, we obtain the (one-dimensional) position operator from quantum me-
chanics. Similarly, for (R3 , B(R3 ), λ) and mj (x) = xj for j = 1, 2, 3 we obtain the
self-adjointness of the momentum operators x̂, ŷ and ẑ in three-dimensional
space. It is now time to treat some more basic operators / observables from
quantum mechanics which are given by differential operators. We start with
momentum operators.
51
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
n o
D(A) = f ∈ L2 (Rn ) : Dj f exists in the weak sense and is in L2 (Rn ) .
∂f
−i = iF −1 Mixj F f = F −1 Mxj F .
∂xj
∂
Moreover, it follows that the unitary operator F maps the domain of −i ∂x
j
bijectively onto the domain of the multiplication operator Mxj . This shows
that A = F −1 Mxj F with the appropriate domain mapping condition D(Mxj ) =
F (D(A)). Since the Fourier transform F on L2 (Rn ) is a unitary operator
∂
by Theorem 2.1.34, we have shown that −i ∂x is unitarily equivalent to the
j
multiplication operator Mxj . Hence, the self-adjointness of the momentum
operators directly follows from the self-adjointness of the multiplication
operators shown in Example 2.2.28.
Notice that in the above example we have in particular considered the op-
d
erator −i dx on L2 (R) with domain H 1 (R) = H01 (R) which is a self-adjoint oper-
ator, whereas for the analogous situation on L2 ([0, ∞)) with domain H01 ([0, ∞))
one does not obtain a self-adjoint operator a shown in Example 2.2.25.
Using the same methods one can treat the Hamiltonian of a free particle
in Rn .
where ∆ denotes the Laplace operator. Of course, we are not done with this
formal calculation because we must also deal with the domain of Ĥ. Let us
52
2.2. Symmetric and Self-Adjoint Operators
∂2
start with the domain of ∂2 xk
. For a function f ∈ L2 (Rn ) to be in its domain
we require that the k-th partial derivative of f exists and again lies in L2 (Rn ).
∂f
Moreover, also the weak k-th partial derivative of ∂x must exist and lie in
k
L2 (Rn ). Taking the Fourier description obtained in Proposition 2.1.35 we see
that xk2 F f ∈ L2 (Rn ) for all k = 1, . . . , n. Equivalently, (1 + |x|2 )F f ∈ L2 (Rn ). As
already used in the proof of the Sobolev embedding theorem (Theorem 2.1.36)
this is equivalent to f ∈ H 2 (Rn ). Hence, Ĥ = −∆ with domain H 2 (Rn ). Arguing
as in the previous example, we see that −∆ is unitarily equivalent to the
multiplication operator M|x|2 via the Fourier transform. Hence, −∆ with
domain H 2 (Rn ) is a self-adjoint operator.
kA∗ k = sup kA∗ f k = sup sup hg|A∗ f i = sup sup hAg|f i = sup sup hAg|f i
kf k≤1 kf k≤1 kgk≤1 kf k≤1 kgk≤1 kgk≤1 kf k≤1
53
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Lemma 2.2.33. Let (A, D(A)) be a densely defined operator on some Hilbert space
H. Then the following are equivalent.
(i) A∗ is injective;
Proof. (ii) ⇒ (i): Let x ∈ D(A∗ ) with A∗ x = 0. Then one has for all y ∈ D(A)
Since A has dense range, we conclude that 0 = hx|zi for all z ∈ H. Choosing
z = x, this shows x = 0. Hence, A∗ is injective.
(i) ⇒ (ii): Assume that A has not dense range. Then the closure Rg A
is a proper closed subspace of H. Then there exists some x ∈ H \ Rg A. By
replacing x by x − P x where P is the orthogonal projection onto RgA, we may
assume that x , 0 and x is orthogonal to RgA. Then we have for all y ∈ D(A)
the identity 0 = hx|Ayi. This shows that x ∈ D(A∗ ) and A∗ x = 0. Since x , 0 by
construction, A∗ cannot be injective which contradicts the assumption.
Theorem 2.2.34. Let (A, D(A)) be a symmetric operator on some Hilbert space H.
Then the following are equivalent.
(iii) Rg(A ± i) = H.
Proof. (i) ⇒ (ii): If A is self-adjoint, one has A = A∗ . Since the latter is closed
by Proposition 2.2.24, A is a closed operator. Moreover, one has Ker(A∗ ± i) =
Ker(A ± i) = 0, since the latter is injective by Lemma 2.2.12.
(ii) ⇒ (iii): By Lemma 2.2.33 the operators A ± i have dense range. We now
show that this already implies that A ± i are surjective. We only treat the case
of A + i as the other case works completely analogously. Recall that one has
54
2.2. Symmetric and Self-Adjoint Operators
the estimate k(A + i)xk ≥ kxk for all x ∈ D(A) by Lemma 2.2.12. Now let y ∈ H.
Since Rg(A + i) is dense, there exist yn ∈ H and xn ∈ D(A) with (A + i)xn = yn
and yn → y in H. One has
Since the right hand side converges it is Cauchy sequence. By the above
estimate (xn )n∈N is a Cauchy sequence as well. Hence, there exists x ∈ H with
xn → x. To summarize we have xn → x and (A + i)xn → y. Since A is closed,
this implies x ∈ D(A) and (A + i)x = y. This shows that A + i is surjective.
(iii) ⇒ (i): Note first that by Lemma 2.2.14 the domain D(A) is dense. It
therefore makes sense to consider the adjoint. By Lemma 2.2.26, the adjoint
A∗ is an extension of A. It therefore remains to show that A extends A∗ . For
this let y ∈ D(A∗ ). By assumption, there exists x ∈ D(A) with (A + i)x = (A∗ + i)y.
Since A∗ extends A, we have (A∗ + i)x = (A∗ + i)y. It follows from Lemma 2.2.33
that (A − i)∗ = A∗ + i is injective. Hence, y = x ∈ D(A) and A extends A∗ .
We now show using the example of real multiplication operators that this
result simplifies verifying self-adjointness.
Note the above argument works without checking the denseness of the
domain or determining the adjoint. With a similar condition one can also
check whether a symmetric operator is essentially self-adjoint or not.
Theorem 2.2.36. Let (A, D(A)) by a densely defined symmetric operator on some
Hilbert space H. Then the following are equivalent.
(ii) Ker(A∗ ± i) = 0;
55
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
We will only give a sketch of the proof of the theorem here and leave
the details as an exercise to the reader. One can show that for a symmetric
operator A∗∗ is again symmetric and the closure of A. Moreover, one always
has A∗ = A∗∗∗ . Hence, the above result follows from Theorem 2.2.34 applied to
A∗∗ and Lemma 2.2.33.
Remark 2.2.37. Recall that in Example 2.2.25 we have seen a closed densely
defined symmetric operator A whose adjoint A∗ is not symmetric. In fact,
this is no coincidence. Suppose that A is a closed densely defined symmetric
operator whose adjoint A∗ is also symmetric. If follows from the symmetry of
A that A ⊂ A∗ . Analogously, it follows from the symmetry of A∗ that A∗ ⊂ A∗∗ .
Since, A is closed, it follows from the comments before this remark that
A = A∗∗ . Hence, A∗ ⊂ A and therefore A is self-adjoint.
Note that as after Definition 2.1.33 it follows from the polarization identity
that (U , D(U )) is an isometry if and only if kU xk = kxk for all x ∈ D(U ). As a
motivation for the following argument we take a quick look at the Möbius
transform
z−i
z 7→ .
z+i
It is a well-known fact from complex analysis that this defines a bijection
between the real line and the unit circle in the complex plane without the
56
2.2. Symmetric and Self-Adjoint Operators
1+w
w 7→ i .
1−w
We now apply the above transformation to a symmetric operator. This trans-
formation is called the Cayley transformation. In the next proposition we study
its basic properties.
V B (A − i)(A + i)−1
This again shows that A ± i is injective and therefore that the inverse (A + i)−1
is well-defined as an unbounded operator with domain Rg(A + i). Moreover,
for x ∈ D(A) we have
k(A + i)xk2 = k(A − i)xk2 .
Now, it follows for x ∈ Rg(A + i) and y = (A + i)−1 x that
Lemma 2.2.40. Let (A, D(A)) be a symmetric operator on some Hilbert space H.
Then the Cayley transform of A is unitary on H if and only if A is densely defined
and self-adjoint.
Proof. First assume that A is self-adjoint. Then it follows from Theorem 2.2.34
that Rg(A ± i) = H. By Proposition 2.2.39 the Cayley transform V is an
isometric operator from H = Rg(A + i) onto H = Rg(A − i). This yields that V is
a surjective isometric operator H → H. It now follows from the polarization
identity that V is unitary on H.
Conversely, if V : H → H is unitary, then Rg(A ± i) = H by Proposi-
tion 2.2.39. The self-adjointness now follows from Theorem 2.2.34.
57
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Lemma 2.2.41. Let H be a Hilbert space and V ∈ B(H) a unitary operator such
that Id −V is injective. Then there exists a self-adjoint operator (A, D(A)) on H
such that V is the Cayley transform of A.
Now, one can verify with a direct computation that the Cayley transform of A
is indeed V , a task which is left to the reader.
Note that the last step of the above proof only used the fact that V is
isometric. Hence, the proof shows that for every isometry U there exists a
symmetric operator A whose Cayley transform is U .
By the above arguments the problem of finding self-adjoint extensions of a
symmetric operator A is reduced to the problem of finding unitary extensions
of the Cayley transform V : Rg(A + i) → Rg(A − i). Now working with orthog-
onal complements becomes quite handy, a notion which we already have used
implicitly several times.
58
2.2. Symmetric and Self-Adjoint Operators
Lemma 2.2.43. Let (A, D(A)) be a densely defined unbounded operator on some
Hilbert space H. Then
Ker A∗ = Rg(A)⊥ .
Hence, x ∈ Rg(A)⊥ . Conversely, let x ∈ Rg(A)⊥ . Then we have for all y ∈ D(A)
the identity hx|Ayi = 0. This implies that x ∈ D(A∗ ) and A∗ x = 0. Therefore
x ∈ Ker A∗ .
We now return to the above extension problem. Note that if such a unitary
extension U : H → H exists, then U splits as a componentwise isomorphism
Hence, Rg(A + i)⊥ and Rg(A + i)⊥ are isomorphic as Hilbert spaces. Conversely,
one sees directly that an isometry Rg(A + i) → Rg(A − i) extends uniquely to an
isometry between the closures. Hence, a unitary isomorphism between Rg(A +
i)⊥ and Rg(A − i)⊥ yields a unitary operator U : H → H extending the Cayley
transform of A. One can check that for such an extension U the operator Id −U
is injective and therefore is the Cayley transform of some self-adjoint operator
B by Lemma 2.2.41. In fact, this gives an one-to-correspondence between
unitary extensions of the Cayley transform and self-adjoint extensions of A. A
crucial role is played by the dimensions of the orthogonal complements just
used.
Definition 2.2.44 (Deficiency indices). The deficiency indices d+ (A) and d− (A)
of a densely defined symmetric operator (A, D(A)) are defined as
Note that by Lemma 2.2.43 one has d+ (A) = dim Rg(A − i)⊥ and d− (A) =
dim Rg(A + i)⊥ . We have just sketched a proof of the following theorem (we
omit a detailed proof).
Theorem 2.2.45. Let (A, D(A)) be a densely defined symmetric operator on some
Hilbert space H. Then
59
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
In fact, more generally one can show that there is an one-to-one correspon-
dence between isometries (U , D(U )) into Ker(A∗ + i) with D(U ) ⊂ Ker(A∗ − i)
and closed symmetric extensions of A. Let us illustrate the theorem with some
examples.
As above all solutions of the above equation are scalar multiples of f (t) = e−t
which are square-integrable. Hence, d− (A) = 1. Altogether we have d+ (A) ,
d− (A) and we see from Theorem 2.2.45 that A has no self-adjoint extensions.
60
2.2. Symmetric and Self-Adjoint Operators
(A∗ + i)f = 0 ⇔ f 0 − f = 0.
Moreover, one can directly verify or use the theory of Möbius transformations
that z 7→ (e − z)/(1 − ez) restricts to a bijection on the unit circle. Hence, there
61
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
exists a β ∈ [0, 2π) (uniquely determined by α) such that for every f ∈ D(Uα )
we have f (0) = eiβ f (1). Conversely, one sees that every f ∈ H 1 ((0, 1)) with
f (0) = eiβ f (1) can be written as g + λ(f− Uα f− ) for some g ∈ H01 ((0, 1)) and some
λ ∈ C. Hence, the self-adjoint operators (Aβ , D(Aβ )) defined by
Given a compactly supported wave function ψ ∈ Cc∞ ((0, 1)) the time evolution
described by the momentum operator should be given by translation of ψ. But
this of course does not completely describe the time evolution for all times
because after some time the wave function will meet the boundary. Since the
time evolution must be described by some unitary operator, we cannot loose
probability mass of the function. Hence, what goes out at one side of the well
should come in at the other side. However, we have the freedom to choose a
phase shift for the outcoming wave. By the superposition principle, this phase
shift is independent of the wave function. In this way one exactly obtains the
just calculated self-adjoint extensions, where eiβ is the global phase shift of
the outcoming waves.
What can we learn from this example? Different self-adjoint extensions
really correspond to different physics! Hence, the concrete choice of a self-
adjoint extension of a given symmetric operator is not just some mathematical
freedom which is irrelevant for the description of the physical world, but has
real physical consequences.
62
2.2. Symmetric and Self-Adjoint Operators
This shows that the second derivative of f exists in the weak sense and is
given by −A∗ f . Conversely, if the second derivative of f exists in the weak
sense and lies in L2 ((0, ∞)), the same calculations read backwards show that
f ∈ D(A∗ ) and A∗ f = −f 00 . Now let us determine the deficiency indices of A.
For d+ (A) we solve
(A∗ + i)f = −f 00 + if = 0.
It follows from the elliptic regularity theory presented in Example 2.1.38 (the
half-line case follows from the real line case via a reduction argument, simply
extend all solutions to the real line by reflecting the solution along at the
origin; in fact this argument also shows D(A∗ ) = H 2 ((0, ∞)) that a solution of
the equation in the weak sense already satisfies f ∈ C ∞ ((0, ∞)) and therefore is
a classical solution. Every classical solution of this second ordinary differential
equation is of the form
√ √
λ1 exp((1 − i)x/ 2) + λ2 exp(−(1 + i)x/ 2)
ik − a
ψ 0 (0) + aψ(0) = −ik + ikλ + a + λa = 0 ⇔ λ= .
ik + a
63
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
(A∗ + i)f = 0 ⇔ f (x) − 2x3 f 0 (x) − 3x2 f (x) = (1 − 3x2 )f (x) − 2x3 f 0 (x) = 0.
We can restrict the above equation to the open right and left half-plane (this
is also valid for distributions). For x > 0 respectively x < 0 we obtain the
64
2.2. Symmetric and Self-Adjoint Operators
equation
1 − 3x2
f 0 (x) = f (x).
2x3
We now treat the case x > 0. If f ∈ L2 (R) solves the equation on [ε, ∞) for
some ε > 0, then it follows from the fact the factor on the right hand side is
bounded that f 0 ∈ L2 ([ε, ∞)). Hence, f ∈ H 1 ([ε, ∞)) because we will see from
the definition of distributions that if the distributional derivative exists as a
locally integrable function, then the weak derivative exists and agrees with the
function representing the distributional derivative. Iterating this argument
and using the Sobolev embeddings we see that f ∈ C ∞ ((ε, ∞)). Hence, f is
smooth away from zero. Hence, away from zero it is sufficient to work with
classical solution of the differential equation. Using separation of variables,
we see that all solutions are scalar multiples of
1 − 3y 2 exp(−1/4y 2 )
Z !
1 −2 3
f+ (x) = exp dy = exp − y − log y = .
2y 3 4 2 y 3/2
Analogously, one obtains for the case x < 0 that the solution is a multiple of
1 − 3y 2 exp(−1/4y 2 )
Z !
1 −2 3
f− (x) = exp dy = exp − y − log|y| = .
2y 3 4 2 |y|3/2
One now sees directly that both are square-integrable in the respective half-
planes. Hence, λ1 f+ 1[0,∞) + λ2 f− 1(−∞,0] ∈ Ker(A∗ + i) for λ1 , λ2 ∈ C. This shows
d+ (A) = 2.
Let us continue with d− (A). For this we have to find all solutions of
(A∗ − i)f = 0 ⇔ f (x) + 2x3 f 0 (x) + 3x2 f (x) = (1 + 3x2 )f (x) + 2x3 f (x) = 0.
65
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
66
2.2. Symmetric and Self-Adjoint Operators
∞
X ∞
X
kgk2 ≤ cn−1/2 2−n k1An k2 = cn−1/2 2−n cn1/2 = 1.
n=1 n=1
Let f (ω) = g(ω)/(z − m(ω)) be the unique solution of the resolvent equation.
Then for all n ∈ N we have
This shows that f is not square integrable and that z − Mm is not surjective. A
fortiori, z ∈ σ (A) and the converse inclusion essim(A) ⊂ σ (A) is shown.
Example 2.2.52. For j = 1, . . . , n let xˆj and pˆj be the position and momentum
operators on Rn as introduced before and in Example 2.2.29. These operators
are self-adjoint and satisfy
Note that the full spectrum of the position and momentum operators
is exactly the result we would expect from physics: if one measures the
momentum (or the position) of a particle, one can obtain arbitrary real values
as measurement outcomes.
An analogous approach yields the spectrum of the negative Laplacian −∆
with domain H 2 (Rn ).
67
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Note that the result σ (−∆) = [0, ∞) perfectly corresponds to the physical
intuition. In fact, recall that −∆ = nj=1 p̂j2 is the Hamiltonian for the free
P
(i) For every sequence (xn )n∈N ⊂ K there exists a subsequence (xnk )k∈N and
x ∈ K with xnk → x for k → ∞.
68
2.3. The Spectral Theorem for Self-Adjoint Operators
The second point is the definition of compact subsets for general topo-
logical spaces, whereas the (in the case of metric spaces equivalent) second
point is called sequentially compactness. You have learned in calculus that a
subset K of some finite dimensional normed vector space V is compact if and
only if K is closed and bounded. This changes fundamentally in the infinite
dimensional setting.
Example 2.3.2 (Non-compact unit ball). Let B be the closed unit ball of some
infinite dimensional separable Hilbert space H. Then B clearly is bounded
and closed by definition. However, we now show that B is not compact.
Choose an orthonormal basis (en )n∈N of H which exists by Theorem 2.1.7. By
orthogonality we have for n , m
The reader should verify the following elementary facts: a compact linear
operator T : H1 → H2 is automatically bounded and T (B) is compact for
arbitrary bounded subsets B of H1 . Moreover, the compact operators form
a subspace of B(H1 , H2 ) and the composition of a bounded and a compact
operator is again a compact operator (hence, in algebraic terms, K(H) is an
ideal in B(H)).
We start our journey through the spectral theorems with the easiest (infi-
nite dimensional) case, namely the spectral theorem for compact self-adjoint
or more generally normal operators.
69
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Theorem 2.3.5 (The spectral theorem for compact normal operators). Let
T : H → H be a compact normal operator on some Hilbert space H. Then there
exists a countable orthonormal system (en )n∈N and (λn )n∈N ⊂ C \ {0} such that
H = Ker T ⊕ span{en : n ∈ N}
and X
Tx = λn hen |xien for all x ∈ H.
n∈N
Moreover, the only possible accumulation point of (λn )n∈N is 0 and every non-zero
eigenvalue of T has finite multiplicity, i.e. dim Ker(T − λ) < ∞ for all λ , 0 and
all eigenvalues λn are real if T is self-adjoint.
Corollary 2.3.6. Let (A, D(A)) be a self-adjoint operator on some separable infinite-
dimensional Hilbert space H with compact resolvent, i.e. R(λ, A) is a compact
operator for some λ in the resolvent set ρ(A). Then there exists an orthonormal
basis (en )n∈N of H and a sequence (λn )n∈N ⊂ R with limn→∞ |λn | = ∞ such that
X∞
2 2
D(A) = x ∈ H : |λ | |he |xi| < ∞
n n
n=1
∞
X
Ax = λn hen |xien .
n=1
Proof. Consider the bounded operator R(λ, A) = (λ − A)−1 . One can verify
that its adjoint is given by R(λ, A∗ ) = R(λ, A). It is easy to see that the re-
solvents at different values commute, in particular we have R(λ, A)R(λ, A) =
R(λ, A)R(λ, A). Hence, R(λ, A) is a compact normal operator. By the spec-
tral theorem for such operators (Theorem 2.3.5) there exists an orthonormal
system (en )n∈N of H and sequence (µn )n∈N with limn→∞ µn = 0 such that
∞
X
R(λ, A) = µn hen |xien .
n=1
Since R(λ, A) is invertible, its kernel is trivial and therefore (en )n∈N must be
an orthonormal basis of H. For λn = λ − µ−1 n define the operator B as in the
70
2.3. The Spectral Theorem for Self-Adjoint Operators
Note that by definition of B we have λ ∈ ρ(B) and R(λ, A) = R(λ, B). In other
words, (λ − A)−1 = (λ − B)−1 . Hence, λ − A = λ − B as unbounded operators and
therefore A = B. Further observe that since A = B is self-adjoint, all eigenvalues
λn of A must be real. Further limn→∞ |µn | = 0 implies limn→∞ |λn | = ∞.
We now give a physical example that illustrate the above methods. How-
ever, before we present a powerful characterization of compact subsets of
Lp -spaces. Its proof uses a smooth approximation of the members of F for
which the Arzelà–Ascoli theorem is applied.
Then the closure of F|Ω is compact in Lp (Ω) for any measurable subset Ω ⊂ Rn of
finite measure. Moreover F is compact in Lp (Rn ) if one additionally has
Z
lim |f (x)|p dx = 0 uniformly on F .
r→∞ |x|≥r
71
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
We have f˜ ∈ H 1 (R) for all f ∈ F . Moreover, verify with the help of Re-
mark 2.1.37 that there exists a universal constant C > 0 such that kf˜kH 1 (R) ≤
C kf kH 1 ((a,b)) . We set F˜ = {f˜ : f ∈ F }. Note that F˜|(a,b) = F . We now check that
F˜ satisfies the condition of the Fréchet–Kolmogorov theorem (Theorem 2.3.7).
First assume that f ∈ C 1 (R)∩ F˜ . Then the Cauchy–Schwarz inequality implies
Z x+h Z x+h !1/2
1/2
|f (x + h) − f (x)| ≤ 0
|f (y)| dy ≤ |h| 0 2
|f (y)| dy ≤ |h|1/2 kf 0 k22
x x
1/2
≤ |h| kf kH 1 (R) .
Now let f˜ ∈ F˜ be arbitrary. Then there exists a sequence (fn )n∈N ⊂ C 1 (R) ∩ F˜
with fn → f˜ ∈ H 1 (R). In particular we have
From this inequality we immediately get that the left hand side converges
uniformly to zero as h → 0. This shows that the closure of bounded sets in
H 1 ((a, b)) are compact subsets of L2 ((a, b)). This result can also be rephrased
in the following way: the natural inclusion
72
2.3. The Spectral Theorem for Self-Adjoint Operators
the Arzelà–Ascoli theorem gives the stronger fact that the Sobolev embedding
H 1 ((a, b)) ,→ C([a, b]) is compact.
We now apply the above embedding to a particle in an infinitely high
square well.
for some −∞ < a < b < ∞. Ignoring physical constants, the Hamiltonian of this
2
system is a self-adjoint extension of A = − dd2 x with D(A) = Cc∞ ((a, b)) and the
underlying Hilbert space H = L2 ((a, b)). The most common choice here is
for which we have already solved the eigenvalue problem in the physics parts
of the lecture. We leave it to the reader the verify the self-adjointness of the
operator with this domain. However, there are also other possbile self-adjoint
extensions corresponding to different physics. Let (B, D(B)) be an arbitrary
self-adjoint extension of A. Then we have D(B) ⊂ H 2 ((a, b)). In particular, we
have
R(i, B)L2 ((a, b)) ⊂ D(B) ⊂ H 2 ((a, b)) ⊂ H 1 ((a, b)).
and therefore R(i, B) is compact as the composition of the bounded opera-
tor R(i, B) : L2 ((a, b)) → H 1 ((a, b)) and the compact operator ι : H 1 ((a, b)) →
L2 ((a, b)). Hence, it follows from Corollary 2.3.6 that L2 ((a, b)) has an orthonor-
mal basis consisting of eigenvectors of B.
73
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
74
2.3. The Spectral Theorem for Self-Adjoint Operators
One can show that the space of trace class operators endowed with the
norm k·k1 is a Banach space in which the finite rank operators are dense.
Note that every linear operator on a finite dimensional Hilbert space is
trace class and its trace coincides with the usual trace for matrices. Using
the general functional calculus for self-adjoint operators, we will soon see
that we can give sense to the square of A∗ A for arbitrary bounded operators.
One can then define trace class operators without assuming the compactness
of A. Nevertheless one obtains the same class of operators as one can show
that trace class implies compactness. Moreover, observe that if a trace-class
operator A has a basis of eigenvectors (en )n∈N with corresponding eigenvalues
(λn )n∈N , we obtain
∞
X ∞
X
Tr A = hAen |en i = λn .
n=1 n=1
Note that it makes no sense to consider unbounded operators in connection
with the trace. Even if an unbounded operator A has an orthonormal basis
of eigenvectors, the associated eigenvalues are not summable unless they
are bounded (they must even form a zero sequence) and therefore define a
bounded (even compact) operator.
One now can precisely define a mixed state as a positive trace class oper-
ator ρ ∈ K(H) with Tr ρ = 1. Here positive means that ρ is self-adjoint and
all eigenvalues of ρ are non-negative. It follows from the spectral theorem
for compact self-adjoint operators (Theorem 2.3.5) that there exists an or-
thonormal basis (ψn )n∈N of eigenvectors with eigenvalues pn ≥ 0 such that
ρ(ϕ) = ∞
P
n=1 pn hψn |ϕiψn for all ϕ ∈ H. Hence,
∞
X
Tr ρ = pn = 1.
n=1
75
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
ρ = |ψihψ|,
i.e. as the one-dimensional projection onto the span of |ψi (provided |ψi is
normalized).
Remark 2.3.12. Please be careful: there are two different notions in quantum
mechanics related with probabilistic measurements. Let us illustrate the two
concepts in a two-dimensional Hilbert space C2 with pure states ψ1 = (1, 0)T
and ψ2 = (0, 1)T which are both eigenvalues for the observable Ĥ = diag(1, 2).
Both states are eigenstates for H and therefore have a certain measurement
outcome, namely 1 for ψ1 and 2 for ψ2 . Now consider the superposition
1
ψ = √ (ψ1 + ψ2 ).
2
This is again a pure state. Measuring the observable Ĥ, we will obtain both
outcomes 1 and 2 with probability equal to 1. Nevertheless we will know for
sure that the system is in the state ψ. Now consider the density matrix
!
1 T T 1/2 0
ρ = (ψ1 ψ1 + ψ2 ψ2 ) = .
2 0 1/2
76
2.3. The Spectral Theorem for Self-Adjoint Operators
Remark 2.3.13. If you still feel unsure about pure and mixed states, the fol-
lowing analogy to classical mechanics may help: in classical mechanics the
state of a system (say particle) is determined by its coordinate in the phase
space, e.g. by its position and momentum, whereas in analogy a quantum
mechanical systems is determined by its state ψ ∈ H. The time evolution of
these states is then governed by the Hamiltonian respectively Schrödinger
equations. In practice, one however often has to work with huge ensembles of
individual particles. In classical mechanics this is usually seen as a collection
of individual systems in different states and modelled via a probability dis-
tribution over the phase space. The quantum mechanical analogue for such
ensembles are mixed states which are represented by densities operators.
We now state some basic properties of trace class operators which are
commonly used in practice.
77
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
(a) They form a subspace of B(H) and the composition of a bounded and a trace
class operator is again trace class.
(b) The trace mapping from the trace class operators is linear and satisfies
Tr(ST ) = Tr(T S)
M : ` 2 (N) → ` 2 (N)
(xn )n∈N 7→ (λn xn )n∈N
Recall that we have seen in Example 2.2.28 that Mm with the above domain
is a self-adjoint operator. The spectral theorem shows that up to unitary
equivalence every self-adjoint operator is of this form. The spectral theorem is
78
2.3. The Spectral Theorem for Self-Adjoint Operators
usually first proved for bounded normal operators. The case of an unbounded
self-adjoint operator A is then reduced to the bounded normal case via the
Cayley transform of A. However, we do not want to delve into the details and
refer to the literature instead. Instead, we present an example.
Example 2.3.17 (The spectral theorem for −∆). We have seen that −∆ with
domain H 2 (Rn ) is a self-adjoint operator in Example 2.2.30. As exploited in
the argument for its self-adjointness, −∆ is equivalent to the multiplication
operator Mm on L2 (Rn ) with m(x) = |x|2 via the unitary operator given by the
Fourier transform F : L2 (Rn ) → L2 (Rn ).
Note that the spectral theorem allows one to define a functional calculus
for self-adjoint operators. In fact, if A is a self-adjoint operator and f : σ (A) →
C is a measurable function, we may define the closed operator
We will discuss this functional calculus in more detail soon when considering
the other variant of the spectral theorem via spectral measures.
Before given the exact definition of such measures, let us again motivative
this version of the spectral theorem. Let A be a self-adjoint operator on Cn .
Let P1 , . . . , Pk denote the orthogonal projections onto the pairwise orthogonal
eigenspaces of A for the eigenvalues λ1 , . . . , λk . Since by the spectral theorem
these eigenspaces span the complete space Cn , we obtain
k k
X X
A = A · I = A
λl Pl =
λ l Pl .
l=1 l=1
Hence, A can be decomposed as the sum of the orthogonal projections onto its
eigenspaces. The natural generalization of the family (P1 , . . . , Pk ) is the concept
of a projection-valued measure.
[ ! N
X
P Ωn x = lim P (Ωn )x for all x ∈ H.
N →∞
n∈N n=1
79
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
80
2.3. The Spectral Theorem for Self-Adjoint Operators
Note that the above properties imply that for all x ∈ H the resolution of
the function λ 7→ hx|Pλ xi is a distribution function of the bounded measure
Ω 7→ hx|P (Ω)xi. If kxk = 1, this is a distribution function of a probability
measure. Conversely, a standard result from measure and probability theory
says that for every bounded distribution function there exist a unique bounded
measure with the given distribution function. This allows one to recover the
spectral measure P from its set of distribution functions λ 7→ hx|Pλ xi for x ∈ H.
For general x, y ∈ H we moreover obtain (complex-valued!) measures related
to the just considered distribution functions via the polarization identity by
4
1X k
hy|P (Ω)xi = i hx + i k y|P (Ω)(x + i k y)i
4
k=1
for Ω ∈ B(R).
Note that for x ∈ H the map Ω 7→ hx|P (Ω)xi defines a Borel measure in the
usual sense of measure theory because of
Hence, we can apply the usual theory of Lebesuge integration to these mea-
sures. In particular, a measurable function f : R → K is said to be finite
almost everywhere with respect to P if it is finite almost everywhere with
respect to all measures hx|P xi for x ∈ H.
We now can state the spectral theorem in its version for projection-valued
measures and its various consequences. To motivate the domains involved
below, consider a step function f = nk=1 ak 1Ωk for some pairwise disjoint
P
Ωk ∈ B(R) and ak ∈ C. Then as for the usual Lebesgue integral we can directly
define the bounded operator
Z n
X
f (λ) dP (λ) B ak P (Ωk ).
R k=1
Note that for the value of the norm of this operator evaluated at some x ∈ H
we obtain
Z n
2 *X n
X +
f (λ) dP (λ)x
= ak P (Ωk )x al P (Ωl )x
R k=1 l=1
n X
X n Xn X
n
= ak al hP (Ωk )x|P (Ωl )xi = ak ak hx|P (Ωk )P (Ωl )xi
k=1 l=1 k=1 l=1
Xn X n Xn
= ak ak hx|P (Ωk ∩ Ωl )xi = |ak |2 hx|P (Ωk )xi
k=1 l=1 k=1
81
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Z
= |f (λ)|2 dhx|P (λ)xi.
R
The above isometry allows one to extend the integral on the left hand side
pointwise for a given x ∈ H for all measurable functions f : R → C which are
square-integrable with respect to the measure Ω 7→ hx|P (Ω)xi = kP (Ω)xk2 via
a limiting argument. Moreover, this calculation together with the Cauchy
Schwarz inequality also shows the less exact inequality
Z
f (λ) dhy|P (λ)xi ≤
kf k∞ kxk kyk
R
(a) We have ( Z )
2
D(A) = x ∈ H : λ dhx|P (λ)xi < ∞
R
and for every x ∈ D(A) and y ∈ H we have
Z
hy|Axi = λ dhy|P (λ)xi.
R
82
2.3. The Spectral Theorem for Self-Adjoint Operators
83
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Z n
X Z n
X Z
= ak 1m−1 (Ωk ) dµ = ak 1Ωk ◦ m dµ = f ◦ m dµ
A∩B k=1 A∩B k=1 A∩B
Z Z
= 1A∩B f ◦ m dµ = 1A 1B f ◦ m dµ.
X X
Using the monotone convergence theorem we see that the above identity
extends to all positive functions f : R → R. By linearity one then can pass to
all functions which are integrable with respect to the measure Ω 7→ hu|P (Ω)wi.
In particular, the identity holds for all bounded functions if A and B have
finite measure. Since both sides of the equality are sesquilinear, the identity
therefore holds for all integrable step functions u and w and all bounded
f . Since the step functions are dense in L2 (X, Σ, µ), the identity extends to
all square integrable functions u and w. Written this out, we have for all
u, w ∈ L2 (X, Σ, µ) and all measurable bounded f : R → R that
Z Z
f (λ) dhu|P (λ)wi = uwf ◦ m dµ = hu|Mf ◦m wiL2 (X,Σ,µ) .
R X
In fact, notice that this example actually gives a proof of the spectral
theorem in the version with projection-valued measures based on the mul-
tiplicative version of the spectral theorem. In fact, one can verify that one
version of the spectral theorem implies the other. The arguments for the con-
verse implication can be found in [Tes12, Lemma 3.3 and following results].
Example 2.3.22 (PVM for orthonormal bases of eigenvectors). Let (A, D(A))
be a self-adjoint operator on a separable infinite-dimensional Hilbert space
and (en )n∈N an orthonormal basis of eigenvectors for A for the real eigenvalues
(λn )n∈N . For Ω ∈ B(R) we define
X
P (Ω) = |en ihen |.
n:λn ∈Ω
Then P (Ω) is the orthogonal projection onto the span of the eigenspaces for
the eigenvalues which lie in Ω. Moreover, P is a projection-valued measure.
In fact, after an unitary transformation this example is again a special case of
the previous one. Nevertheless let us verify directly the required properties.
In fact, everything is clear except for the σ -additivity. For this note that one a
direct decomposition
M
H= Hk ,
k
84
2.3. The Spectral Theorem for Self-Adjoint Operators
where Hk are the pairwise orthogonal eigenspaces for the different eigen-
values λk of A. Let Pk be the orthogonal projection onto Hk . Then Pk =
P
n:λn =λk |en ihen |. It follows from orthogonality that
X
kxk2 = kPk xk2 for all x ∈ H. (2.2)
k
Now, let (Ωn )n∈N ⊂ B(R) be pairwise disjoint. Then we have for x ∈ H
X N
2
X X
2
P (Ωn )x − P (∪n∈N Ωn )x
=
Pk x − Pk x
n=1 k:λk ∈∪N
n=1 Ωn
k:λk ∈∪∞
n=1 Ωn
X
2 X
=
Pk x
= kPk xk2 −−−−−→ 0.
N →∞
k:λk ∈∪∞
n=N +1 k:λk ∈∪∞
n=N +1
The Born–von Neumann formula We have seen that for quantum mechanics
S is the set of all trace class operators on some fixed (separable) Hilbert
space H and A is the set of all self-adjoint operators on H. Let PA be the
projection-valued measure of A. Then µA is given by the Born–von Neumann
formula
µA (Ω) = Tr PA (Ω)ρ for Ω ∈ B(R).
Note that the trace is well-defined because PA (E)ρ is trace class as the compo-
sition of a bounded and a trace class operator. However, one still has to check
that µA is a probability measure. The only non-trivial fact is the σ -additivity
of µA . For this let (Ωk )k∈N be pairwise disjoint. Then for an orthonormal basis
85
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
(en )n∈N
N
X ∞
X
µA (Ωk ) = Tr(PA (∪N
k=1 Ωk )ρ) = hen |PA (∪nk=1 Ωk )ρen i.
k=1 n=1
Note that for all n ∈ N the sequence PA (∪Nk=1 Ωk )en converges to PA (∪k∈N Ωk )en
by the properties of a projection-valued measure. Since moreover the n-th
summand is dominated by hen |ρen i which is summable because ρ is trace
class, the series converges to Tr(PA (∪k∈N Ωk )ρ) by the dominated convergence
theorem.
Further, the expectation value of the observable A in the state ρ is (pro-
vided it exists) Z
hAiρ = λ dµA (λ).
R
Some useful results We now state some useful related results without
proofs which are often used in physics. The first result can be verified by
simply checking the definitions and shows that under mild assumptions the
expectation value of a state can be calculated as shown in the physics part of
the lecture.
Proposition 2.3.23. Let A be an observable and ρ a state such that hAiρ exists (as
finite value) and Im ρ ⊂ D(A). Then Aρ is trace class and
hAiρ = Tr(Aρ).
One can show that the above definition is equivalent to the fact that the re-
solvents of A and B commute, i.e. one has R(λ, A)R(µ, B) = R(µ, B)R(λ, A) for all
λ, µ ∈ C \ R. For commuting observables one has the following generalization
of the spectral theorem.
86
2.3. The Spectral Theorem for Self-Adjoint Operators
Moreover, as in the case of a single operator one obtains a joint functional calculus
by integrating measurable functions on Rn against the spectral measure.
However, for the above formula to define a probability measure for all states
ρ and all Ω as above we need that Ω1 × · · · × Ωn 7→ PA1 (Ω1 ) · · · PAn (Ωn ) extends
to a projection-valued measure on Rn . Since the product of two orthogonal
projections is an orthogonal projection if and only if the projections commute,
we see that the observables must commute pairwise. From a physical perspec-
tive this agrees with the requirement that the simultaneous measurement of
several observables should be independent of the order of the measurements
of the individual observables.
87
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
∂
i~ ψ(t) = Hψ(t).
∂t
From a mathematical perspective there are several problems with the above
equation. First of all the equation does not make sense in the usual meaning
if the initial value ψ0 or the solution ψ(t) does not lie in the domain of H. So
how one has to interpret the above equation? And even if ψ0 ∈ D(H), why
does the above equation have a unique solution? In fact, if H is unbounded,
the usual theorems for existence and uniqueness of solutions involving a
Lipschitz condition do not apply.
Note that the series is absolute convergent in B(H). In fact, we have for t ∈ R
because of the boundedness of H
∞ ∞
X k(iHt)k k X kHkk |t|k
≤ = e|t|kHk .
k! k!
k=0 k=0
88
2.3. The Spectral Theorem for Self-Adjoint Operators
89
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Hence, U (t)x0 ∈ D(A) and x(t) is differentiable with ẋ(t) = −iAU (t)x0 =
−iU (t)Ax0 . Now, let y be a second solution of the problem. For t > 0 consider
the function z(s) = U (t − s)y(s). Then it follows from a variant of the product
rule that
This shows that ż(s) = 0 for all s ∈ R. By reducing to the case of the functions
s 7→ hw|z(s)i for w ∈ H, we see from the scalar case that the function z is
constant. In particular, we have U (t)x0 = z(0) = z(t) = y(t). This establishes
the uniqueness of the solutions.
90
2.3. The Spectral Theorem for Self-Adjoint Operators
Hence, for elements in the domain of the generator we obtain classical solu-
tions of the Schrödinger equation. In the general case x0 ∈ H we may therefore
interpret U (t)x0 as a generalized solution of the Schrödinger equation.
As one would hope the infinitesimal generator of every strongly continu-
ous unitary group is self-adjoint. As a preliminary results in this direction we
show the following lemma.
Hence, xR ∈ D(A) and (iA − 1)xR = e−R U (R)x − x. Note that both xR and
(iA − 1)xR convergeR as R → ∞. Since A is closed this implies that the improper
∞
Riemann integral 0 e−t U (t)x dt lies in the domain of A and
Z∞
(iA − 1) e−t U (t)x dt = lim e−R U (R)x − x = −x.
0 R→∞
This shows that (U ∗ (t))t∈R is a unitary group. We now show that the group is
strongly continuous. For this observe that for all x, y ∈ H
91
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Since (U ∗ (t))
t∈R is strongly continuous, we may calculate its infinitesimal
generator B. Note that because of
U ∗ (h)x − x U (−h)x − x
−i = −i
h h
one has B = −A. By the first part of the proof we know that −A + i and
therefore also A − i are surjective. The self-adjointness of A now follows from
Theorem 2.2.34. But this is by definition equivalent to A being essentially
self-adjoint.
We now come to the fundamental theorem of this section which shows that
the description of the evolution via the Schrödinger equation is equivalent to
the description via unitary groups.
A 7→ (eitA )t∈R .
92
2.3. The Spectral Theorem for Self-Adjoint Operators
Proof. Let (U (t))t∈R be a strongly continuous unitary group. Then its in-
finitesimal generator A is essentially self-adjoint by Lemma 2.3.30. Therefore
the self-adjoint operator A generates the strongly continuous unitary group
(eitA )t∈R . Since A is self-adjoint, we see that D(A) and therefore also D(A)
are dense in H. Note that for x ∈ D(A) both unitary groups yield classical
solutions for the problem ẋ(t) = Ax(t) with initial value x0 . By the unique-
ness of the solutions shown in Proposition 2.3.29, the operators eitA and
U (t) therefore agree on the dense subset D(A) for all t ∈ R. Since these op-
erators are bounded, we indeed have eitA = U (t) for all t ∈ R. This shows
(eitA )t∈R = (U (t))t∈R . In particular, as the groups agree, so do the generators
and we have A = A. Altogether we have shown that the map A 7→ (eitA )t∈R is
onto. For the injectivity simply note again that if (eitA )t∈R and (eitB )t∈R define
the same unitary groups, then one clearly has A = B.
We now study the unitary groups generated by the position and momen-
tum operators. Since both are particular instances of multiplication operators,
we study this class first.
Corollary 2.3.34. Let n ∈ N and for j = 1, . . . , n let x̂j denote the position operator
on L2 (Rn ) in direction j. Then x̂j generates the unitary group
Hence, these groups simply act on pure states by a uniform phase rotation.
After a Fourier transform an analogous argument applies to the momentum
operators.
∂
Corollary 2.3.35. Let n ∈ N and for j = 1, . . . , n let p̂j = −i ∂x be the momentum
j
93
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Proof. Recall that we have seen in Example 2.2.29 that under the unitary
Fourier transform p̂j becomes the multiplication operator with xj . Hence, it
follows that for all f ∈ L2 (Rn ) one has almost everywhere
Requiring linearity of the time evolution for mixed states, the same formula
holds for (non-normalized) mixed states made of finitely many pure states.
Assuming again continuity of the time evolution, it follows from the density of
the finite rank operators in the space of trace class operators that the identity
indeed holds for all trace class operators. In particular, it holds for all density
operators. Hence, we obtain the following mathematical description of the
time evolution in quantum mechanics.
ρ 7→ U (t)ρU (t)−1 .
−∆ + V .
94
2.4. Further Criteria for Self-Adjointness
Theorem 2.4.2 (von Neumann’s criterion). Let (A, D(A)) be a densely defined
symmetric operator on a Hilbert space H. If there is a conjugation C : H → H such
that C(D(A)) ⊂ D(A) and
AC = CA on D(A),
Proof. We first show that C(D(A∗ )) ⊂ D(A∗ ) and that A∗ C = CA∗ . For this let
x ∈ D(A) and y ∈ D(A∗ ). Then by assumption
95
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Here we have used integration by parts twice for the first summand. Hence,
−∆ + V is a densely defined symmetric operator on L2 (Rn ). Now observe that
the complex conjugation C : f 7→ f clearly is a conjugation that leaves Cc∞ (Rn )
invariant and statisfies AC = CA on Cc∞ (Rn ). Hence, by von Neumann’s
criterion (Theorem 2.4.2) the operator A has self-adjoint extensions.
Theorem 2.4.4 (Kato–Rellich). Let (A, D(A)) and (B, D(B)) be two self-adjoint
operators on a Hilbert space H. Suppose that D(A) ⊂ D(B) and that there exist
constants 0 < a < 1 and 0 < b such that
Proof. Use a positive sufficiently large number µ ∈ R such that a + b/µ < 1.
This is possible because of a < 1. Now for all x ∈ D(A) we have
96
2.4. Further Criteria for Self-Adjointness
The second factor on the right hand side is invertible because A is self-adjoint.
Now let us deal with the first factor. Observe that because of D(A) ⊂ D(B) this
factor is a closed operator which is defined on the whole Hilbert space. For all
x ∈ H we have by the assumption and the estimate of Lemma 2.2.12 that
kxk2 = k(A(A + iµ)−1 x + iµ(A + iµ)−1 xk2 = k(A(A + iµ)−1 xk2 + kµ(A + iµ)−1 xk2
+ 2 RehA(A + iµ)−1 x|iµ(A + iµ)−1 xi = k(A(A + iµ)−1 xk2 + kµ(A + iµ)−1 xk2 .
Forgetting the second summand at the right hand side, we therefore obtain
Since the factor on the right is smaller than 1 by the choice of µ, using the
Neumann series (Lemma 2.2.20) we see that B(A + iµ)−1 + Id and therefore
A + B + iµ = µ(µ−1 A + µ−1 B + i) are invertible. Of course, the same argument
applies to (A + B − iµ). It now follows from Theorem 2.2.34 that A + B is
self-adjoint with domain D(A).
kf k∞ ≤ kF f k1 ≤ Cε (ε k∆f k2 + kf k2 ).
97
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
Hence, for all ε > 0 there exists a constant cε > 0 such that
kf k∞ ≤ ε k∆f k2 + cε kf k2 .
One can show that the theorem still holds for n ≥ 4 if one replaces the
assumption V1 ∈ L2 (Rn ) by V1 ∈ Lp (Rn ) for some p > n/2. For a proof see
[RS75, Theorem X.20]. As a particular instance of the above theorem we
obtain the self-adjointness of the hydrogen atom.
1 1
V (x) = V1 (x) + V2 (x) = 1|x|≤1 + 1|x|>1 .
|x| |x|
1
Hence, it follows from Theorem 2.4.5 that −∆ + |x| is self-adjoint with domain
2 3
H (R ).
With more effort one can prove a variant of Theorem 2.4.5 due to T. Kato
which allows a wider range of potentials. However, in this case one only
obtains essentially self-adjointness on the space of test functions and one has
no exact information on the domain of the closure. We omit the proof because
of its complexity and refer to [RS75, Theorem X.29].
98
2.4. Further Criteria for Self-Adjointness
The theorem again holds for n ≥ 4 if one replaces the condition V1 ∈ L2 (Rn )
with V1 ∈ Lp (Rn ) for some p > n/2. Note that the above theorem in particular
applies for non-negative continuous potentials. For example, −∆ + x4 is
essentially self-adjoint on Cc∞ (Rn ). The sign of the potential here plays a
crucial role as the next example makes clear. We do not discuss it here and
refer the reader to [Hal13, Section 9.10] where a detailed exposition is given.
Example 2.4.8. The operator −∆ − x4 with domain Cc∞ (Rn ) is not essentially
self-adjoint.
In contrast to this negative result recall that we have seen in Example 2.4.3
that −∆ − x4 has self-adjoint extensions. In fact, since −∆ − x4 on Cc∞ (Rn ) is not
essentially self-adjoint, the above example even yields that there must exist
several self-adjoint extensions.
Observe that the defining condition for an analytic vector implies that the
kAn xk n
map z 7→ ∞
P
n=0 n! z is well-defined and analytic. In particular, it follows
form the fact that derivatives are analytic and the triangle inequality that
span{An x : n ∈ N} entirely consists of analytic vectors if x is analytic.
Let us consider a self-adjoint operator (A, D(A)) on H with the projection-
valued measure P . Then for M > 0 the operator A leaves P ([−M, M])H invari-
ant and therefore restricts to a bounded operator on P ([−M, M])H. Therefore
An n
the exponential series ∞
P
n=0 n! t converges in absolutely operator norm for
99
2. The Theory of Self-Adjoint Operators in Hilbert Spaces
all t ∈ R. This implies that for x in the dense set ∪M∈N P ([−M, M])H we have
a C ∞ -vector
∞
X kAn xk n
t <∞ for all t > 0.
n!
n=0
Hence, for a self-adjoint operators its set of analytic vector is dense. In fact,
this property characterizes self-adjoint operators. This is Nelson’s criterion
for self-adjointness which we will not prove here due to time constraints. A
proof can be found in [Mor13, Theorem 5.47] and [RS75, Theorem X.39].
Example 2.4.11 (Harmonic oscillator). Recall from the physics part that
d2 2
the quantum mechanical harmonic oscillator is given by − dx 2 + x ignoring
physical constants. This time the space of test functions Cc∞ (R) is not the
optimal choice as we have already seen in the physics part that the operator
has Hermite functions as eigenvectors. Therefore we work with the space of
Schwartz function S(R) instead. This is is the space of all C ∞ -functions with
rapid decay at infinity. This space plays a fundamental role in the theory of
distributions where you can also find an exact definition (Definition 3.2.1).
d2 2 with domain S(R) is a
A direct calculation now shows that − dx 2 + x
d d
a=x− and a† = x +
dx dx
the Hermite polynomials. Note that we have used the notation a† instead of
a∗ because a† is not the adjoint of a. Moreover, there are no domain problems
because for arbitrary polynomial expressions in a and ↠their domain is again
S(R) if a and a† are chosen with domain S(R) because S(R) is left invariant
by those operators.
We now sketch how one can show that (ψn )n∈N in fact forms an orthonor-
mal basis of L2 (R). Therefore we must show that V = span{ψn : n ∈ N} is
dense in L2 (R). Note that it follows from the fact that the polynomials pn
100
2.4. Further Criteria for Self-Adjointness
2
have degree n that V = {p(x)e−x /2 : p polynomial}. One now shows that for all
α∈C
N
X α n xn −x2 /2 2
e −−−−−→ eαx e−x /2 in L2 (R).
n! n→∞
n=0
Hence, if ψ is orthogornal to every element in the closure of V , we have
Z
2
e−ikx e−x /2 ψ(x) dx = 0 for all k ∈ R.
R
2
This means that the Fourier transform of e−x /2 ψ(x) is identically zero. By
2
Plancherel’s theorem (Theorem 2.1.34) this implies that e−x /2 ψ(x) and there-
⊥
fore also ψ(x) are the zero function. Hence, V = V ⊥ = 0 which is equivalent
to the denseness of V in L2 (Rn ).
Now observe that eigenvalues of an operator A are clearly analytic vectors
d2 2
for A. Hence, (ψn )n∈N is a set of analytic vectors for − dx 2 + x whose span is
2
d
dense in L2 (Rn ). Nelson’s criterion (Theorem 2.4.10) now shows that − dx 2 +x
2
101
Distributions 3
We have already seen that the domains of self-adjoint realizations of differen-
tial operators contain functions which are not classically differentiable, i.e.
non-differentiable Sobolev functions. However, there are still many locally in-
tegrable functions which do not have a derivative even in the weak sense. This
disadvantage has already made some arguments very difficult or impossible,
for example in Example 2.1.38 or Example 2.2.49. In this section we give an
introduction to the theory of distributions which also can be differentiated and
have the huge advantage that they are closed with respect to differentiation.
Moreover, we have formally seen (see Example 2.2.5) that such generalized
functions can naturally arise as some kind of generalized eigenfunctions for
self-adjoint operators which may not have eigenvalues in the usual sense. In
fact, the spectral theorem can be generalized to this situation and we will see
that typical quantum mechanical operators have an orthonormal basis if one
allows for eigenvectors in this generalized sense. This is the statement of the
main results of this chapter, the so-called nuclear spectral theorem.
(i) there exists a compact subset K ⊂ Ω such that supp fn ⊂ K for all n ∈ N
and
We call Cc (Ω) with this notation of convergence the space of all test functions
on Ω und write D(Ω).
103
3. Distributions
More precisely, there exists a locally convex topology on the vector space
Cc (Ω) for which the convergence of a sequence (fn )n∈N is equivalent to (i) and
(ii). However, we will ignore such topological concepts to a great extent and
instead work directly with the notion of convergence induced by the topology.
Distributions are then defined as continuous linear functionals on D(Ω).
Let us start with the most prominent example of all, the infamous Dirac
distribution.
δa : D(Ω) → C
ϕ 7→ ϕ(a).
One has the following useful criterion for a linear functional u : D(Ω) → C
to be a distribution.
104
3.1. The Space of Distributions
Along the same line one can identify finite Borel measures with distribu-
tions.
One can show that the embedding µ 7→ uµ from the locally finite Borel mea-
sures into the space of distributions is injective. This can be seen as a stronger
105
3. Distributions
Remark 3.1.9 (For experts). The notation D0 (Ω) for the space of distributions
actually has a deeper mathematical meaning. It stands for the topological dual
of the space D(Ω), i.e. the space of all continuous linear functionals on D(Ω).
The above definition says that a sequence of distributions converges in D0 (Ω) if
and only it converges with respect to the weak∗ -topology. However, for deeper
results on distributions it may be more suitable to not endow D0 (Ω) with
the weak∗ -topology. Indeed, often D0 (Ω) is defined with a different topology
(the so-called strong topology) which however agrees on bounded sets (note
that we have and will not define the notion of boundedness on topological
vector spaces) with the weak∗ -topology. Hence, the notion of convergence of
sequences is independent of the choice between these topologies.
for all n ∈ N. Suppose further that for all ε > 0 one has
Z
lim fn (x) dx = 0.
n→∞ |x|≥ε
106
3.1. The Space of Distributions
This means that the masses of the functions fn concentrate on arbitrary small
intervals as n goes to infinity. Under these assumptions one has fn → δ0 in
D0 (Ω). Indeed, for ϕ ∈ D(Ω) one has
Z∞ Z ∞
|hδ0 , ϕi − hfn , ϕi| = ϕ(0) − fn (x)ϕ(x) dx = fn (x)(ϕ(0) − ϕ(x)) dx
−∞ ∞
Z∞
≤ fn (x) |ϕ(0) − ϕ(x)| dx
−∞
n 1 n
fn (x) = √ exp(−(nx)2 ) and gn (x) =
π π 1 + (nx)2
or using the substitution ε(n) = n−1/2 one obtains the approximations for
ε = ε(n) → 0
x2
!
1 1 ε
fε (x) = √ exp − and gε (x) = .
πε ε π ε2 + x2
107
3. Distributions
Example 3.1.12 (Derivative of the delta function). Consider the Dirac dis-
tribution δ0 ∈ D0 (R). For ϕ ∈ D(Ω) and n ∈ N one has
∂n ϕ
hD n δ0 , ϕi = (−1)n hδ0 , D n ϕi = (−1)n (0).
∂n x
Hence, the distributional derivative of the Dirac distribution is given by (−1)n
times the evaluation of the n-th derivative at zero.
Hence, the distributional derivative of sign is given by two times the Dirac
distribution.
108
3.2. Tempered Distributions
Note that if ϕ ∈ D(Ω) and m is smooth, then the product m · ϕ is again a test
function. This allows us to extend the above identity as a definition to general
distributions.
Note that one sees directly that m·u is linear and continuous and therefore
a well-defined distribution. Let us give a very elementary example to see how
one works with such multiplications.
We then define the Schwartz space or the space of all rapidly decreasing functions
as
S(Rn ) B {f ∈ C ∞ (Rn ) : kf kα,β < ∞ for all α, β ∈ Nn0 }.
109
3. Distributions
More intuitively, the above definition means that f and all of its derivatives
exist and decay faster than any inverse power of x. As for the test functions
D(Ω) we need a notion of convergence for Schwartz functions.
Definition 3.2.2 (Convergence of Schwartz functions). We say that a se-
quence of Schwartz functions (fm )m∈N ⊂ S(Rn ) converges to f ∈ S(Rn ) if
kfm − f kα,β → 0 for all α, β ∈ Nn0 .
It is a good exercise to check that (fm )m∈N ⊂ S(Rn ) tends to f ∈ S(Rn ) as
m → ∞ if and only if (fm )m∈N converges to f with respect to the metric
X kf − gkα,β
d(f , g) B 2−(|α|+|β |) (f , g ∈ S(Rn )).
1 + kf − gkα,β
α,β∈Nn0
Moreover, one can show that (S(Rn ), d) is a complete metric space. These
properties show in mathematical terms that S(Rn ) is a so-called Fréchet space.
Let us give the exact definition for later use. First we need the concept of a
semi-norm which we have already met in our study of Lp -spaces.
Definition 3.2.3 (Semi-norm). Let V be a K-vector space. A map p : V →
[0, ∞) is called a semi-norm on V if
(i) p is homogeneous, i.e. p(λx) = λp(x) for all x ∈ V and λ ∈ K;
(ii) p satisfies the triangle inequality, i.e. p(x + y) ≤ p(x) + p(y) for all x, y ∈ V .
Hence, the only difference between a norm and a semi-norm is the fact
that a semi-norm p may not satisfy the definiteness condition p(x) = 0 ⇔ x = 0.
Fréchet spaces are defined in terms of a countable family of semi-norms.
Definition 3.2.4 (Fréchet space). Let G be a vector space and (pk )k∈N a count-
able family of semi-norms on G with the following properties:
(i) If x ∈ G satisfies pk (x) = 0 for all k ∈ N, then x = 0.
(ii) If (xn )n∈N ⊂ G is Cauchy with respect to each semi-norm pk , i.e. for all
k ∈ N and for all ε > 0 there exists N ∈ N such that
pk (xn − xm ) < ε for all n, m ≥ N ,
then there exists x ∈ G such that (xn )n∈N converges to x with respect to
each semi-norm, i.e. pk (xn − x) → 0 as n → ∞ for all k ∈ N.
Then G together with the datum of these semi-norms is called a Fréchet space.
In this case (G, d) is a complete metric space for the metric
∞
X kf − gkk
d(f , g) B 2−k (f , g ∈ G).
1 + kf − gkk
k=1
110
3.2. Tempered Distributions
Remark 3.2.5. One can always replace the family (pk )k∈N by a family of
increasing semi-norms (qk )k∈N by setting qk = supn≤k qn . In fact, the family
(qk )k∈N induces an equivalent metric on G and we are only interested in the
topological properties and not on the concrete values of the metric itself.
Hence, we may assume that the semi-norms satisfy
Note that in contrast to this situation for S(Rn ) one can show that the
convergence in D(Ω) for ∅ , Ω ⊂ Rn is not induced by a translation-invariant
metric.
As for distributions we define the tempered distributions as the space of
all continuous functionals on S(Rn ).
111
3. Distributions
This follows from the fact that under the Fourier transform differentiation
becomes multiplication. The details should be verified by the reader. Note
that moreover for Schwartz functions f , ϕ ∈ S(Rn ) we have by Plancherel’s
formula (Theorem 2.1.34)
Z Z Z
huF f , ϕi = (F f )(x)ϕ(x) dx = −1
f (x)(F ϕ)(x) dx = f (x)(F ϕ)(x) dx
Rn Rn Rn
= huf , F ϕi.
This equality can now be used to extend the Fourier transform to tempered
distributions.
Proof. As in the case of the Fourier transform, one has a well-defined mapping
F −1 : S 0 (Rn ) → S 0 (Rn ). We now show that as the notation already indicates
that F −1 is the inverse of F . In fact, we have for u ∈ S 0 (Rn ) and ϕ ∈ S(Rn )
using the mere definitions
This shows that F −1 F = IdS 0 (Rn ) . An analogous calculation also shows that
F F −1 = IdS 0 (Rn ) .
Let us now give some examples which are particularly relevant for physics.
112
3.2. Tempered Distributions
Hence, we have shown that F δ0 = (2π)−n/2 1. Physically this means that one
obtains the delta function if one takes all frequencies with equal (normalized)
strength.
Hence, we have shown that F (eik· ) = (2π)n/2 δk . Physically this gives the
obvious fact that one obtains a plain wave if one picks a single frequency (via
a delta-distribution).
Lu = δ0 .
113
3. Distributions
hu ∗ g, ϕi = hu, g̃ ∗ ϕi.
114
3.2. Tempered Distributions
We now show for the example of the three dimensional Laplace operator
how a fundamental solution for a linear differential operator with constant
coefficients can in principle be obtained.
This shows that x 7→ |x|−2 ∈ L1loc (R3 ). Therefore its Fourier transform exists in
the sense of tempered distributions. Moreover, one has |x|−2 1B(0,R) → |x|−2 in
S 0 (R3 ) as R → ∞. This is a direct consequence of the dominated convergence
theorem. By the continuity of the Fourier transform, we therefore obtain
F −1 |x|−2 1B(0,R) → F −1 |x|−2 as R → ∞. Hence for ϕ ∈ S(Rn ), we can determine
the action of u on ϕ via the calculation
Let us again first take a look at the inner integral. The following argument
holds for all r > 0 and |y| , 0: Choose an orthogonal matrix O such that Oez =
115
3. Distributions
y/|y|. Then using the substitution s = θz = cos ϕ (in spherical coordinates with
dϕ
√
ϕ ∈ [0, π)) and consequently ds = (arccos s)0 = −1/ 1 − s2 , we obtain
Z Z Z
cos(rθy) dθ = cos(r|y|θ · Oez ) dθ = cos(r|y|O−1 θ · ez ) dθ
S2 Z S 2
Z S2
= cos(r|y|θ · ez ) dθ = cos(r|y|θz ) dθ
S2 S2
Z 1 Z Z 1
ds
=− √ dα cos(r|y|s) √ = −2π cos(r|y|s) ds
−1 1−s2 S1 1 − s2 −1
" #s=1
sin(r|y|s) sin(r|y|)
= −2π = −4π
r|y| s=−1 r|y|
R∞
Recall the identity 0 sinx x dx = π/2 for the sinc-function. By the substitution
formula we therefore obtain for |y| , 0
Z∞
sin(r|y|) π
dr = .
0 r|y| 2|y|
Here we can exchange the limit and the integral with the help of the dominated
convergence theorem, where we use the fact that there exists a universal
constant M > 0 such that
Z Z
R sin x R sin ax M
dx ≤ M for all R > 0 ⇒ dx ≤ for all a, R > 0.
x 0 ax a
0
By a similar calculation one can see that an analogue formula holds for
the case n ≥ 3, whereas for n = 2 one obtains a logarithmic term.
116
3.3. The Nuclear Spectral Theorem
Definition 3.3.1 (Dual of Fréchet spaces). Let G be a Fréchet space. Its dual
space G 0 is defined as
We endow G 0 with the weak∗ -topology, i.e. a sequence (ϕn )n∈N ⊂ G 0 satisfies
ϕn → ϕ ∈ G 0 if and only if ϕn (g) → ϕ(g) for all g ∈ G.
i † : x 7→ [g 7→ hx|i(g)i].
117
3. Distributions
that xn → x in H implies i † (xn )(g) → i † (x)(g) for all g ∈ G. This shows the
continuity of the map i † : H → G 0 . Furthermore, the injectivity of i † follows
from the density of G in H because an element in the dual space H∗ is uniquely
by its values on a dense subset.
i i†
G ,− → G0
→ H ,−
In the following we will often ignore the embedding i and directly identify
elements of G with elements in H. However, one has to be careful when one
identifies elements of H with elements of G 0 because the embedding i † is
anti-linear. The next example is probably the most important Gelfand triple.
i i†
Example 3.3.3. Consider S(Rn ) ,− → L2 (Rn ) ,−
→ S 0 (Rn ), where the embedding
n 2
i : S(R ) → L (R ) is the natural inclusion and i † : H → S 0 (Rn ) is given by
n
" Z #
† n
i : f 7→ S(R ) 3 ϕ 7→ f (x)ϕ(x) dx .
Rn
Hence, except for a change with the complex conjugate i † is the restriction of
the inclusion of L1loc -functions into the space of distributions.
A0 : G 0 → G 0
ϕ 7→ [g 7→ ϕ(Ag) C hϕ, AgiG0 ,G ].
118
3.3. The Nuclear Spectral Theorem
Of course, E(λ)0 may be trivial for certain values of λ ∈ R. One can show
that A has a complete system of generalized eigenvectors if and only if the
span of all E(λ) is dense in G 0 for the weak∗ -topology. As before we will ignore
this topological issues and will instead consider some examples. First we
consider the momentum operator in one dimension.
hA0 u, ϕiS 0 (R),S(R) = hu, AϕiS 0 (R),S(R) = hu, −iϕ 0 iS 0 (R),S(R) = hu 0 , iϕiS 0 (R),S(R)
= hiu 0 , ϕiS 0 (R),S(R) .
119
3. Distributions
This shows that A0 u = iu 0 in the sense of distributions. Note that for ϕ ∈ S(R)
one has A0 i † ϕ = iϕ 0 = i † (−iϕ 0 ) = i † Aϕ as it should hold in general. Now let us
determine the generalized real eigenvalues of A, i.e. the eigenvalues of A0 . For
this we must find for given λ ∈ R all distributional solutions u ∈ S 0 (R) of
A0 u = iu 0 = λu.
One can show that all distributional solutions of the above equation already
are smooth functions (we have already used and proved an easier variant of
this result where we additionally assumed that u 0 ∈ L1loc (R)). Hence, every
solution of the above equation is of the form u(x) = ce−iλx for some c ∈ C.
In other words, this shows that E(λ) = {ce−iλx : c ∈ C} for all λ ∈ R. We now
determine the generalized Fourier transform with respect to A. For ϕ ∈ S(R)
we have Z
ce−iλx 7→ hce−iλx , ϕiS 0 (R),S(R) = c ϕ(x)e−iλx dx.
R
hMx0 u, ϕiS 0 (R),S(R) = hu, Mx ϕiS 0 (R),S(R) = hu, x · ϕiS 0 (R),S(R) = hx · u, ϕiS 0 (R),S(R)
Hence, Mx0 : u 7→ x·u is the multiplication of the function x 7→ x with the given
distribution. Now, Mx0 δx0 = xδx0 = x0 δx0 holds in a mathematical rigorous way.
Moreover, one can show that all solutions of the eigenvalue equation
Mx0 u = x · u = x0 u ⇔ (x − x0 ) · u = 0
are constant multiples of δx0 . In fact, passing to the Fourier transform (for
tempered distributions) the problem reduces to the eigenvalue problem of the
previous example. Put differently, we have E(λ) = {cδλ : c ∈ C} for all λ ∈ R.
120
3.3. The Nuclear Spectral Theorem
Let us now calculate the generalized Fourier transform with respect to A. For
ϕ = cδλ ∈ E(λ) and ϕ ∈ S(R) we have
this identification, for ϕ ∈ S(R) the Fourier transform ϕ̂ agrees with the
function λ 7→ ϕ(λ), that is ϕ̂ = ϕ and the Fourier transform is the identity
mapping. Of course, from this it follows immediately that A has a complete
system of generalized eigenvectors.
The nuclear spectral theorem says that one can always find a complete
system of generalized eigenvalues for a self-adjoint operator A which is com-
patible with a Gelfand triple (G, H, ι) provided the Fréchet space G is nice
enough. This niceness condition is in mathematical terms that G is a so-called
nuclear space. This is a rather abstract mathematical concept from the the-
ory of locally convex topological vector spaces and needs some work to be
properly presented. Physics students may ignore the following definitions
and simply work with the fact that the space of Schwartz functions S(Rn )
is nuclear if they feel overwhelmed by the amount of definitions and may
therefore directly jump to the statement of the nuclear theorem.
From now on let G be a Fréchet space given by a countable family of
increasing semi-norms (pn )n∈N . For such a semi-norm we can define its local
Banach space.
Definition 3.3.8 (Local Banach space). Let G be a Fréchet space and p one
its defining semi-norms. Then for the null space Np = {x ∈ G : p(x) = 0} define
the quotient vector space
Gp B G/Np = {x + Np : x ∈ G}.
Then (Gp , p) is a normed vector space and its completion Ĝp is called the local
Banach space for p.
121
3. Distributions
Example 3.3.11 (The space of smooth functions). Let us consider the space
G = C ∞ ([0, 1]). We naturally want that a sequence in (fn )n∈N ∈ C ∞ ([0, 1]) con-
verges to some f ∈ C ∞ ([0, 1]) if and only if D k fn → D k f uniformly on [0, 1] for
all k ∈ N. This is achieved if we require that (fn )n∈N converges with respect
to all semi-norms pk (f ) B supx∈[0,1] |f (x)| for k ∈ N. One sees that the family
(pk )k∈N satisfies all requirements in the definition of a Fréchet spaces. Hence,
(pk )k∈N gives C ∞ ([0, 1]) the structure of a Fréchet space. Alternatively, one can
also work with the norms qk (f ) = kf kH k (0,1) for k ∈ N. Note that pk (fn − f ) → 0
for all n ∈ N if and only if qk (fn − f ) → 0 for all n ∈ N. This is a consequence
of the fact that by the Sobolev embedding theorems (Theorem 2.1.36) conver-
gence in H m ((0, 1)) implies convergence in C k (Rn ) provided m is large enough.
Observe that the family (qk )k∈N is monotone, i.e qk ≤ qk+1 for all k ∈ N.
We now show that C ∞ ([0, 1]) is a nuclear space. Since qk is already a norm,
the local Banach spaces are given by Gk = H k ((0, 1)). For l > k the natural
map Ĝl → Ĝk is the natural inclusion H l ((0, 1)) ,→ H k ((0, 1)) of Sobolev spaces.
We must show that for l large enough these inclusions are nuclear mappings.
For this let us consider the discrete Fourier transform F : L2 ([0, 1]) → ` 2 (Z)
which is an isomorphism because the trigonometric system (e2πim· )m∈Z is an
orthonormal basis of L2 ([0, 1]). Analogous to the continuous case, the image
of H k ((0, 1)) under F is the space of all sequences (xn )n∈N for which (nk xn )n∈N
lies in ` 2 (Z) or equivalently (xn )n∈N lies in ` 2 (Z, (n2k )). Hence, under the
122
3.3. The Nuclear Spectral Theorem
Fourier transform the inclusion H l ((0, 1)) ,→ H k ((0, 1)) becomes the identity
mapping
` 2 (Z, (n2l )) ,→ ` 2 (Z, (n2k )).
Choose yn = n−2k en , where en is the n-th unit vector, and xn = n−2l en . Then
kxn k`2 (Z,(n2l )) = kyn k`2 (Z,(n2k )) = 1 and (xn ) and (yn ) form orthonormal bases in
the respective spaces. Hence, for all z ∈ ` 2 (Z, (n2l )) we have
∞
X ∞
X
z= hxn |zi`2 (Z,(n2l )) xn = hxn |zi`2 (Z,(n2l )) n2(k−l) yn .
n=1 n=1
This shows that for l > k the identity mapping between the weighted spaces is
nuclear because in this case the sequence (n2(k−l) )n∈N is absolutely summable.
Hence, C ∞ ((0, 1)) is a nuclear Fréchet space.
With the above result one can further easily deduce that the Fréchet space
C ∞ (R)is nuclear. The ambitious reader may try to prove that the space of
Schwartz functions S(Rn ) is a nuclear Fréchet space as well, a fact which
we will use soon. The less ambitious readers can find a proof of this fact in
[Kab14, p. 279].
We now come to the final result of our lecture, the nuclear spectral theo-
rem.
and XZ
Ax = λhηλ,k , xiηλ,k dµk (λ)
k∈K R
We now sketch some main steps of the proof. For a complete proof see
[Kab14, Satz 16.41]. One first establishes the following refined variant of
the spectral theorem for unbounded self-adjoint operators: for a self-adjoint
operator (A, D(A)) on H there exists a family of finite Borel measures (µk )k∈K
123
3. Distributions
This version of the spectral theorem can first be proven for bounded normal
operators and thereafter the self-adjoint case can be deduced with the help
of the Cayley transform. An elegant way to prove the spectral theorem for
a normal operator T ∈ B(H) is to establish a continuous functional calculus
C(σ (T )) → B(H) for T . One then decomposes the above representation into
the direct sum of cyclic representations which can be shown to be unitary
equivalent to multiplication operators. For proofs of these facts we refer to
the literature on functional analysis given in the bibliography.
With the above unitary transform one can essentially reduce the prob-
lem to the case of multiplication operators. One now tries to define delta-
distributions δk,λ for k ∈ K and λ ∈ R. These distributions would be general-
ized eigenfunctions for A provided they are well-defined elements in G 0 . In
order to prove that these delta-distributions are indeed well-defined one uses
the fact that G is nuclear. In fact, one can show that if K is a nuclear Fréchet
space and one has a Gelfand triple of the form K ,→ L2 (Ω, Σ, µ) ,→ K0 for some
measure space (Ω, Σ, µ), then for almost all ω ∈ Ω the delta-distributions δω
are well-defined elements of K0 .
As a particular instance of the nuclear spectral theorem we obtain the
following corollary for differential operators which includes some quantum
mechanical operators.
124
3.3. The Nuclear Spectral Theorem
for such general nuclear locally convex spaces which may not be Fréchet
spaces. We have avoided general locally convex spaces in this lecture to
reduce topological difficulties. In fact, the topology of general locally convex
space is not induced by a (translation-invariant) metric in contrast to the
case of Fréchet spaces and concepts such as continuity are more difficult to
formulate correctly. A prominent and important example of a nuclear space
which is not a Fréchet space is the space of test functions D(Ω) for some open
Ω , ∅. Taking the nuclear spectral theorem for granted in the case of D(Ω),
we obtain the following strengthening of the previous corollary.
Note that the above corollary can in fact be applied to the Hamiltonian
Ĥ = −∆ + |x|−1 of the hydrogen atom. In fact, we can choose Ω = R3 \ {0} as the
potential is smooth outside the singularity at zero. The above corollary then
shows that there exists a complete system of generalized eigenfunctions in
D0 (R3 \ {0}).
125
The postulates of quantum A
mechanics
In this short appendix we present all postulates of quantum mechanics used
in our lectures in a condensed and mathematical manner. Our formulation of
the postulates closely follows the presentation in [Tak08].
127
A. The postulates of quantum mechanics
128
Bibliography
[Bar95] Robert G. Bartle. The elements of integration and Lebesgue measure.
Wiley Classics Library, vol. Containing a corrected reprint of the
1966 original [ıt The elements of integration, Wiley, New York;
MR0200398 (34 #293)], A Wiley-Interscience Publication. John
Wiley & Sons, Inc., New York, 1995, pp. xii+179.
[Bog07] V. I. Bogachev. Measure theory. Vol. I, II. Springer-Verlag, Berlin,
2007, Vol. I: xviii+500 pp., Vol. II: xiv+575.
[Hal13] Brian C. Hall. Quantum theory for mathematicians. Graduate Texts
in Mathematics, vol. 267. Springer, New York, 2013, pp. xvi+554.
[Kab14] Winfried Kaballo. Aufbaukurs Funktionalanalysis und Operatortheo-
rie. Springer-Verlag, 2014.
[Mor13] Valter Moretti. Spectral theory and quantum mechanics. Unitext,
vol. 64. With an introduction to the algebraic formulation, La
Matematica per il 3+2. Springer, Milan, 2013, pp. xvi+728.
[RS75] Michael Reed and Barry Simon. Methods of modern mathemati-
cal physics. II. Fourier analysis, self-adjointness. Academic Press
[Harcourt Brace Jovanovich, Publishers], New York-London, 1975,
pp. xv+361.
[Rud87] Walter Rudin. Real and complex analysis. Third. McGraw-Hill Book
Co., New York, 1987, pp. xiv+416.
[Tak08] Leon A. Takhtajan. Quantum mechanics for mathematicians. Gradu-
ate Studies in Mathematics, vol. 95. American Mathematical Society,
Providence, RI, 2008, pp. xvi+387.
[Tes12] Gerald Teschl. Ordinary differential equations and dynamical systems.
Graduate Studies in Mathematics, vol. 140. American Mathematical
Society, Providence, RI, 2012, pp. xii+356.
129