Functional Analysis With Applications
Functional Analysis With Applications
A quote:
Over the decades, Functional Analysis has been enriched and inspired on account of
demands from neighboring fields, within mathematics, harmonic analysis (wavelets
and signal processing), numerical analysis (finite element methods, discretiza-
tion. . . ), PDEs (diffusion equations, scattering theory), representation theory; iter-
ated function systems (fractals, Julia sets, chaotic dynamical systems), ergodic the-
ory, operator algebras, and many more. And neighboring areas, probability/statistics
(for example stochastic processes, Ito and Malliavin calculus), physics (representa-
tion of Lie groups, quantum field theory), and spectral theory for Schrödinger oper-
ators.
5
Preface
7
Acknowledgment
The first named author thanks his students in the Functional Analysis sequence 313-
314. Also he thanks postdocs and colleagues; among them: D. Alpay, W. Arveson,
O. Bratteli, P. Casazza, I. Cho, D. Dutkay, R. Kadison, W. Klink„ K. Kornelson, D.
Larson, P. Muhly, R. Niedzialomski, J. Packer, E. Pearse, R.S. Phillips, W. Polyzou,
D. Robinson, S. Sakai, I.E. Segal, K. Shuman, R. Werner.
9
Contents
1 Elementary Facts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.1 A Sample of Topics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1.2 Duality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
1.3 Transfinite Induction (Zorn and All That) . . . . . . . . . . . . . . . . . . . . . . 27
1.4 Basics of Hilbert Space Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
1.4.1 Orthonormal Bases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
1.4.2 Bounded Operators in Hilbert Space . . . . . . . . . . . . . . . . . . . . 35
1.5 Dirac’s Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
1.5.1 Three Norm-Completions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
1.5.2 Connection to Quantum Mechanics . . . . . . . . . . . . . . . . . . . . . 44
1.5.3 Probabilistic Interpretation of Parseval in Hilbert Space . . . . 48
1.6 The Lattice Structure of Projections . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
11
12 Contents
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 325
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337
Part I
Introduction and Motivation
17
Subject Example
´x
A analysis f (x) − f (0) = 0 f 0 (y) dy
Table 0.1: Examples of Linear Spaces: Banach spaces, Banach algebras, Hilbert
spaces H , linear operators act in H .
Reader Guide.
Below we list chapter by chapter how the six areas in Table 0.1 are covered.
Ch 1: Areas A, E, F.
Ch 2: Areas A, C, F.
Ch 3: Areas B, C, E, F.
Ch 4: Areas A, B, E, F.
Ch 5: Areas A, E, F.
Ch 6: Area E.
Ch 7: Areas D, F.
Ch 8: Areas E, F.
Ch 9: Areas B, C, D.
Ch 10: Areas A, F.
Ch 11: Areas A, B, C, D, E.
In more detail, the six areas in Table 0.1 may be fleshed out as follows:
Problems worthy
of attack
prove their worth
by hitting back.
— Piet Hein
Below we outline some basic concepts, ideas, and examples which will be studied
inside the book itself. While they represent only a sample, and we favor the setting
of Hilbert space, the details below still tie in nicely with diverse tools and techniques
not directly related to Hilbert space.
23
24 1 Elementary Facts
Classical functional analysis is roughly divided into two branches, each with a long
list of subbranches:
PQ − QP = iI (1.1)
The objective is to build a Hilbert space such that the symbols P and Q are rep-
resented by unbounded selfadjoint operators, defined on a common dense domain
in some Hilbert space, and with the operators satisfying (1.1) on this domain. In
class (ii), we consider both C∗ -algebras and von Neumann algebras (also called W ∗ -
algebras); and in case (iii), our focus is on unitary representations of the group G
under consideration. The group may be abelian or non-abelian, continuous or dis-
crete, locally compact or not. Our present focus will be the case when G is a Lie
group. In this case, we will study its representations with the use of the correspond-
ing Lie algebra.
• wavelets theory
1.2 Duality 25
• harmonic analysis
Our notions from analytic number theory will be those that connect to groups,
and representations; such as the study of automorphic forms, and of properties of
generalized zeta-functions.
1.2 Duality
The “functional” in the name “Functional Analysis” derives from the abstract notion
of a linear functional: Let E be a vector space over a field F (we shall take F = R,
or F = C below.)
If E comes with a topology (for example from a norm, or from a system of semi-
norms), we say that ϕ is a continuous linear functional. Occasionally, continuity
will be implicit.
Definition 1.2. The set of all continuous linear functionals is denoted E ∗ , and it is
called the dual space. (In many examples there is a natural identification of E ∗ as
illustrated in the following examples (Table 1.1).
Definition 1.3. If E is a normed vector space, and if it is complete in the given norm,
we say that E is a Banach space.
26 1 Elementary Facts
Proof. An exercise.
E E∗ how?
1
l p , 1 ≤ p < ∞ with lq, p + q1 = 1 x = (xi ) ∈ l p , y = (yi ) ∈ l q ,
l p -norm ϕy (x) = ∑i xi yi
´1
C (I), I = [0, 1] with signed Borel ϕµ ( f ) = 0 f (x) dµ (x), ∀ f ∈ C (I) .
max-norm measures µ on I, of
bounded variation
!1
p
p
kxk p := ∑ |xk | .
k∈N
ˆ 1
p
p
kϕk p = |ϕ (x)| dµ (x) .
R
´ ´
(By Stieltjes integral, |ϕ| p dµ = |ϕ| p dF.)
Remark 1.2. We gave the definition of L p (µ) in the case where µ = dF, but it ap-
plies more generally.
Theorem 1.1 (Banach). Let E be a Banach space, and let x ∈ E\ {0}, then there is
a ϕ ∈ E ∗ such that ϕ (x) = kxk.
Remark 1.3. In the examples (i) and (ii) above, i.e., l p or L p (µ), it is possible to
identify the needed elements in E ∗ . But the power of Theorem 1.1 is that it yields
existence for all Banach spaces, i.e., when E is given only by the axioms from
Definitions 1.1-1.2.
Let (X, ≤) be a partially ordered set. By partial ordering, we mean a binary relation
“≤” on the set X, such that (i) x ≤ x; (ii) x ≤ y and y ≤ x implies x = y; and (iii)
x ≤ y and y ≤ z implies x ≤ z.
A subset C ⊂ X is said to be a chain, or totally ordered, if x, y ∈ C implies that
either x ≤ y or y ≤ x. Zorn’s lemma says that if every chain has a majorant then there
exists a maximal element in X.
Theorem 1.2 (Zorn). Let (X, ≤) be a partially ordered set. If every chain C in X
has a majorant (or upper bound), then there exists an element m in X so that x ≥ m
implies x = m.
An illuminating example of a partially ordered set is the binary tree model (Figs
1.1-1.2). Another example is when X is a family of subsets of a given set, partially
ordered by inclusion.
28 1 Elementary Facts
m1 m2
m3
m5 m4
m6
Fig. 1.1: Finite Tree (natural order on the set of vertices). Examples of maximal
elements: m1 , m2 , . . .
Zorn’s lemma lies at the foundation of set theory. It is in fact an axiom and is
equivalent to the axiom of choice, and to Hausdorff’s maximality principle.
Theorem 1.3 (Hausdorff Maximality Principle). Let (X, ≤) be a partially ordered
set, then there exists a maximal totally ordered subset L in X.
The axiom of choice is equivalent to the following statement on infinite product,
which itself is extensively used in functional analysis.
Theorem 1.4 (axiom of choice). Let Aα be a family of nonempty sets indexed by
α ∈ I. Then the infinite Cartesian product
Ω = ∏ Aα = ω : I → ∪α∈I Aα ω (α) ∈ Aα
α∈I
is nonempty.
The point of using the axiom of choice is that, if the index set is uncountable, there
is no way to verify whether (xα ) is in Ω , or not. It is just impossible to check for
each α that xα is contained in Aα .
1.4 Basics of Hilbert Space Theory 29
In case the set is countable, we simply apply the down to earth standard induction.
Note that the standard mathematical induction is equivalent to the Peano’s axiom:
Every nonempty subset of the set of natural number has a unique smallest element.
The power of transfinite induction is that it applies to uncountable sets as well.
In applications, the key of using the transfinite induction is to cook up in a clear
way a partially ordered set, so that the maximum element turns out to be the ob-
ject to be constructed. Examples include Hahn-Banach extension theorem, Krein-
Milman’s theorem on compact convex set, existence of orthonormal bases in Hilbert
space, Tychnoff’s theorem on infinite Cartesian product of compact spaces (follows
immediately from the axiom of choice.)
We will apply transfinite induction (Zorn’s lemma) to show that every infinite di-
mensional Hilbert space has an orthonormal basis (ONB).
Definition 1.5. Let (X, k·k) be a normed space. X is called a Banach space if it is
complete with respect to the induced metric
d (x, y) := kx − yk , x, y ∈ X.
Remark 1.4. The abstract formulation of Hilbert was invented by von Neumann in
1925. It fits precisely with the axioms of quantum mechanics (spectral lines, etc.)
A few years before von Neumann’s formulation, Heisenberg translated Max Born’s
quantum mechanics into mathematics.
Lemma 1.2 (Cauchy-Schwarz). 1 Let (X, h·, ·i) be an inner product space, then
Proof. By the positivity axiom in the definition of an inner product, we see that
* +
2 2 2
∑ ci c j xi , x j = ∑ ci xi , ∑ c j x j ≥ 0, ∀c1 , c2 ∈ C;
i, j=1 i=1 j=1
Corollary 1.1. Let (X, h·, ·i) be an inner product space, then
p
kxk := hx, xi, x ∈ X (1.3)
defines a norm.
Proof. It suffices to check the triangle inequality (Definition 1.6). For all x, y ∈ X,
we have
kx + yk2 = hx + y, x + yi
= kxk2 + kyk2 + 2ℜ {hx, yi}
≤ kxk2 + kyk2 + 2 kxk kyk (by (1.2))
= (kxk + kyk)2
1 Hermann Amandus Schwarz (1843 - 1921), German mathematician, contemporary of Weier-
strass, and known for his work in complex analysis. He is the one in many theorems in books on
analytic functions. There are other two “Schwartz” (with a “t”):
Laurent Schwartz (1915 - 2002), French mathematician, Fields Medal in 1950 for his work of
distribution theory.
Jack Schwartz (1930 - 2009), American mathematician, author of the famous book “Linear
Operators”.
1.4 Basics of Hilbert Space Theory 31
Definition 1.7. An inner product space (X, h·, ·i) is called a Hilbert space if X is
complete with respect to the metric
d (x, y) = kx − yk , x, y ∈ X;
Remark 1.5. An extremely useful method to build Hilbert spaces is the GNS con-
struction. For details, see chapter 4. The idea is to start with a positive definite
function ϕ : X × X → C, defined on an arbitrary set X. We say ϕ is positive definite,
if for all n ∈ N,
n
∑ ci c j ϕ (xi , x j ) ≥ 0 (1.4)
i, j=1
∑ cx δx , ∑ dy δy ϕ
:= ∑ cx dy ϕ (x, y) .
Note that
2
∑ cx δx ϕ
:= ∑ cx δx , ∑ cx δx ϕ
= ∑ cx cy ϕ (x, y) ≥ 0
x,y
and set H := completion of the quotient space H0 /N with respect to k·kϕ . (The
fact that N is really a subspace follows from (1.2).) H is a Hilbert space.
Φ : X → H such that
ϕ (x, y) = hΦ (x) , Φ (y)iH (1.5)
for all (x, y) ∈ X × X, where h·, ·iH denotes the inner product in H .
Given a solution Φ satisfying (1.5), then we say that H is minimal if
Proof. These conclusions follow from Remark 1.5, and the definitions. (The miss-
ing details are left as an exercise to the student.)
Remark 1.6. It is possible to be more explicit about choice of the pair (Φ, H ) in
Corollary 1.2, where ϕ : X × X → C is a given positive definite function: We may
in fact choose H to be L2 (Ω , F, P) where P = Pϕ depends on ϕ, and (Ω , F, P) is
a probability space.
see Fig 1.3. In this case, we may then take Ω = C (R) = all continuous functions on
R, and Φt (ω) := ω (t), t ∈ [0, ∞), ω ∈ C (R).
And in this case, further the measure Pϕ on the sigma-algebra in C (R) generated
by cylinder-sets is the Wiener-measure; and Φ on L2 (C (R) , cyl, P) is the standard
Brownian motion, i.e., Φ : [0, ∞) → L2 (C (R) , P) is a Gaussian process with
ˆ
EP (Φ (s) Φ (t)) = Φs (ω) Φt (ω) dP (ω) = s ∧ t.
C(R)
Remark 1.7. We see from the proof of Lemma 1.2 that the Cauchy-Schwarz inequal-
ity holds for all positive definite functions.
1.4 Basics of Hilbert Space Theory 33
s jH×, t)
t
s
s 0 t
(a) ϕ (s,t) = s ∧ t (b) with t fixed
1. uα , uβ H = δαβ and
2. span {uα } = H . (Here “span” means “closure of the linear span.”)
We are ready to prove the existence of orthonormal basis for any Hilbert space.
The key idea is to cook up a partially ordered set satisfying all the requirements in
the transfinite induction, so that the maximum elements turns out to be an orthonor-
mal basis. Notice that all we have at hands are the abstract axioms of a Hilbert space,
and nothing else. Everything will be developed out of these axioms.
Lemma 1.3. Let H be a Hilbert space and S ⊂ H . Then the following are equiv-
alent:
1. x ⊥ S implies x = 0
2. span{S} = H
Proof. Now we prove Theorem 1.6. If H is empty then the proof is done. Other-
wise, let u1 ∈ H . If ku1 k 6= 1, it can be normalized by u1 / ku1 k. Hence we may
assume ku1 k = 1. If span{u1 } = H the proof finishes again, otherwise there exists
/ span{u1 }. By Lemma 1.3, we may assume ku2 k = 1 and u1 ⊥ u2 . By induction,
u2 ∈
there exists a collection S of orthonormal vectors in H .
34 1 Elementary Facts
Let P (S) be the set of all orthonormal sets partially order by set inclusion. Let
C ⊂ P(S) be any chain and let M :=
S
E∈C E. M is clearly a majorant of C.
In fact, M is in the partially ordered system. For if x, y ∈ M, there exist Ex and Ey
in C so that x ∈ Ex and y ∈ Ey . Since C is a chain, we may assume Ex ≤ Ey . Hence
x, y ∈ Ey , and so x ⊥ y. This shows that M is in the partially ordered system and a
majorant.
By Zorn’s lemma, there exists a maximum element m ∈ P (S). It suffices to show
that the closed span of m is H . Suppose this is false, then by Lemma 1.3 there exists
x ∈ H so that x ⊥ span {m}.
Since m ∪ {x} ≥ m and m is maximal, it follows that x ∈ m, which implies x ⊥ x.
By the positivity axiom of the definition of Hilbert space, x = 0.
Uj
scaling: f (x) −−→ 2 j/2 f 2 j x
V k
translation: f (x) −−→ f (x − k)
1.4 Basics of Hilbert Space Theory 35
Lemma 1.4 (Riesz). There is a bijection H 3 h 7−→ lh between H and the space
of all bounded linear functionals on H , where
and kT ∗ k = kT k.
Hence the mapping y 7−→ hx, Tyi is a bounded linear functional on H . By Riesz’s
theorem, there exists a unique hx ∈ H , s.t. hx, Tyi = hhx , yi, for all y ∈ H . Set
T ∗ x := hx . One checks that T ∗ linear, bounded, and in fact kT ∗ k = kT k.
kT ∗ T k = kT k2 . (1.10)
• T is normal if T T ∗ = T ∗ T
• T is selfadjoint if T = T ∗
• T is unitary is T ∗ T = T T ∗ = IH (= the identity operator)
• T is a (selfadjoint) projection if T = T ∗ = T 2
1
R= (T + T ∗ )
2
1
S = (T − T ∗ )
2i
T = R + iS.
This is similar the to decomposition of a complex number into its real and imaginary
parts. Notice also that T is normal if and only if R and S commute. Thus the study
of a family of normal operators is equivalent to the study of a family of commuting
selfadjoint operators.
PH = {x ∈ H : Px = x}
which simplifies to
2
wn + wm
4 x− + kwn − wm k2 = 2 kx − wn k2 + kx − wm k2 . (1.12)
2
H x
W
wn-1
wn
o Px =lim wn
Every Hilbert space has an ONB, but it does not mean in practice it is easy to
select one that works well for a particular problem. The Gram-Schmidt orthogonal-
ization process was developed a little earlier than von Neumann’s formulation of
abstract Hilbert space. It is an important tool to get an orthonormal set out of a set
of linearly independent vectors.
38 1 Elementary Facts
u1
v1 = .
ku1 k
Proof.
x − PFn x
vn+1 := , n = 1, 2, . . . (1.13)
kx − PFn xk
Note the LHS in (1.13) a unit vector, and orthogonal to PFn H .
The formula for PFn is
n
PFn = ∑ |vk ih vk | .
k=1
Remark 1.9. If H is non separable, the standard induction does not work, and the
transfinite induction is needed.
Example 1.3. Let H = L2 [0, 1]. The polynomials {1, x, x2 , . . .} are linearly indepen-
dent in H , for if
n
∑ ck xk = 0
k=1
Vn = span{1, x, . . . , xn−1 }
= span {h0 , h1 , . . . , hn−1 } .
xn+1 − Pn xn+1
hn+1 := , n ∈ N.
kxn+1 − Pn xn+1 k
1.4 Basics of Hilbert Space Theory 39
Example 1.4. Let H = L2 [0, 1]. Consider the set of complex exponentials
i2πnx
e : n ∈ N ∪ {0} ,
is an ONB in L2 (R).
Note with the normalization in (1.14) we get
ˆ ˆ
2
ψ j,k (x) dx = |ψ (x)|2 dx, ∀ j, k ∈ Z.
R R
Example 1.5 (Haar wavelet). Let H = L2 (0, 1), and let ϕ0 be the characteristic
function of the unit interval [0, 1]. ϕ0 is called a scaling function. Define
Proof. Fix k, if j1 6= j2 , then ψ j1 ,k and ψ j2 ,k are orthogonal since their supports are
nested. For fixed j, and k1 6= k2 , then ψ j,k1 and ψ j,k2 have disjoint supports, and so
they are also orthogonal. See Fig 1.5.
Ψj k
2 2
P.A.M Dirac was every efficient with notations, and he introduced the “bra-ket”
vectors [Dir35, Dir47].
Definition 1.12. Let H be a Hilbert space with inner product h·, ·i. We denote by
“bra” for vectors hx| and “ket” for vectors |yi, for x, y ∈ H .
Lemma 1.6. Let v ∈ H be a unit vector. The operator x 7→ hv, xi v can be written as
Pv = |v ih v| ,
Also, if x, y ∈ H then
D E
hx, Pv yi = hx, vi hv, yi = hx, viv, y = hhv, xi v, yi = hPv x, yi
1.5 Dirac’s Notation 41
so Pv = Pv∗ .
PF := ∑ |vi ih vi |
vi ∈F
PF2 =
∑ (|vi ih vi |) v j ih v j = ∑ |vi ih vi | = PF
vi ,v j ∈F vi ∈F
Remark 1.10. More generally, any rank-one operator can be written in Dirac nota-
tion as
|u ih v| : H 3 x 7−→ hv, xi u ∈ H .
With the bra-ket notation, it is easy to verify that the set of rank-one operators forms
an algebra, which easily follows from the fact that
IH = ∑ |uα ih uα | .
α∈J
v= ∑ huα , vi uα , ∀v ∈ H .
α∈J
= ∑ hA∗ ui , uk i uk , Bu j
k
= A∗ ui , Bu j
= ui , ABu j
= (MAB )i j
Exercise 1.4. Let H be a Hilbert space, and A a set which indexes a fixed ONB
{vα }α∈A . Now define T : H → l 2 (A), by
(T h) (α) := hvα , hi , ∀α ∈ A, h ∈ H .
T T ∗ = Il 2 (A) , and
T ∗ T = IH .
• (TN) Set √
kT kT N := trace T ∗T ;
• (HSN) Set
1
kT kHSN = (trace (T ∗ T )) 2 .
Theorem 1.8. The completion of F R (H ) with respect to k·kUN , k·kT N , and k·kHSN
are the compact operators, the trace-class operators, and the Hilbert-Schmidt
operators.
We will prove, as a consequence of the Spectral Theorem (Chapter 3), that the
k·kUN - completion agrees with the usual definition of the compact operators.
Definition 1.14. Let T ∈ B (H ), then we say that T is compact iff (Def.) T (H1 )
is weak∗ -compact in H , where H1 := v ∈ H kvk ≤ 1 .
A similar remark applies to the other two Banach spaces of operators. The
Hilbert-Schmidt operators forms a Hilbert space.
Exercise 1.5. If dim H = ∞, show that the identity operator IH is not compact.
Definition 1.15. The following is useful in working through the arguments above.
Exercise 1.6. Let T ∈ F R (H ), and let the three norms be as in Definition 1.13.
Then show that
kT kUN ≤ kT kHSN ≤ kT kT N ; (1.15)
Remark 1.11. In the literature, the following notation is often used for the three
norms in Definition 1.13:
kT kUN = kT k∞
kT kT N = kT k1 (1.16)
kT k
HSN = kT k2
kT k∞ ≤ kT k2 ≤ kT k1 , T ∈ F R (H ) .
Much of the motivation for the axiomatic approach to the theory of linear opera-
tors in Hilbert space dates back to the early days of quantum mechanics (Planck,
Heisenberg, and Schrödinger), but in the form suggested by J. von Neumann. (von
Neumann’s formulation is the one now adopted by most books on functional anal-
ysis.) Here we will be brief, as a systematic and historical discussion is far beyond
our present scope. Suffice it to mention here that what is known as the "matrix-
mechanics" of Heisenberg takes the form infinite by infinite matrices with entries
representing, in turn, transition probabilities, where "transition" refers to "jumps"
between energy levels. By contrast to matrices, in Schrödinger’s wave mechanics,
the Hilbert space represents wave solutions to Schrödinger’s equation. Now this
entails the study of one-parameter groups of unitary operators in H . In modern
language, with the two settings we get the dichotomy between the case when the
1.5 Dirac’s Notation 45
trace(AB) = trace(BA).
This implies the trace on the left-hand-side is zero, while the trace on the RHS is
not.
This shows that there is no finite dimensional solution to the commutation rela-
tion above, and one is forced to work with infinite dimensional Hilbert space and
operators on it. Notice also that P, Q do not commute, and the above commutation re-
lation leads to the uncertainty principle (Hilbert, Max Born, von Neumann worked
out the mathematics), which says that the statistical variance 4P and 4Q satisfy
4P4Q ≥ h̄/2 . We will come back to this later.
We will show that non-commutativity always yields “uncertainty.”
However, Heisenberg [vN31, Hei69] found his “matrix” solutions by tri-diagonal
∞ × ∞ matrices, where
0 1
√
1 0 2
√ √
1 2 0 3
P= √ (1.18)
2 √ ..
3 0 .
.. ..
. .
and
0 1
√
−1 0 2
√ √
1 − 2 0 3
Q= √ (1.19)
i 2
√ ..
− 3 0 .
.. ..
. .
Exercise 1.7. Using matrix multiplication for ∞ × ∞ matrices verify directly that the
two matrices P and Q satisfy PQ − QP = iI, where I is the identity matrix in l 2 (N0 ),
i.e., (I)i j = δi j . Hint: use the rules in Lemma 1.8.
and
0 0 ··· ··· ··· ···
1 0 0 · · · · · · · · ·
√ 0 √2 0 0
A+ = 2 · · · · · ·
.
.
. √
. 0 3 0 0 · · ·
.. .. √ .. ..
. . 0 4 . .
In other words, the raising operator A+ is a sub-banded matrix, while the lowering
operator A− is a supper-banded matrix. Both A+ and A− has 0s down the diagonal.
Further, show that
1 0 ··· ··· ··· ···
0 2 0 · · · · · · · · ·
.
.
· · · · · ·
A− A+ = 2 . 0 3 0
.. ..
. . 0 4
0 · · ·
.. .. .. .. .. ..
. . . . . .
..
i.e., a diagonal matrix, with the numbers N down the diagonal inside . ; and
0 0 ··· ··· ··· ··· ···
0 1 0 · · · · · · · · · · · ·
..
. 0 2 0 ··· · · · · · ·
A+ A− = 2
...
;
0 3 0 · · · · · ·
. ..
.
. . 0 4 0 · · ·
.. .. .. .. .. ..
. . . . . .
so that
1
[A− , A+ ] = I.
2
48 1 Elementary Facts
Remark 1.13. The conclusion of the discussion above is that the Heisenberg com-
mutation relation (1.17) for pairs of selfadjoint operators has two realizations, one
in L2 (R), and the other in l 2 (N).
1 d
In the first one we have (P f ) (x) = i dx f , and(Q f ) (x) = x f (x), for f ∈ S ⊂
L2 (R), where S denotes the Schwartz test-function subspace in L2 (R).
The second realization is by ∞ × ∞ matrices, and it is given in detail above. In
section 7.3 we shall return to the first realization.
The Stone-von Neumann uniqueness theorem implies the two solutions are uni-
tarily equivalent; see chapter 7.
Case 1. Let H be a complex Hilbert space, and let {uk }k∈N be an ONB, then Par-
seval’s formula reads:
Translating this into a statement about transition probabilities for quantum states,
v, w ∈ H , with kvkH = kwkH = 1, we get
See Fig 1.6. The states v and w are said to be uncorrelated iff (Def.) they are orthog-
onal.
Fix a state w ∈ H , then
1.5 Dirac’s Notation 49
The numbers |hui , wi|2 represent a probability distribution over the index set, where
|huk , wi|2 is the probability that the quantum system is in the state |uk i.
∑ P (A j ) = P (∪ j A j )
j
where A j ∈ B (R), Ai ∩ A j = 0,
/ i 6= j.
3. P (A ∩ B) = P (A) P (B), ∀A, B ∈ B (R).
50 1 Elementary Facts
A∩B χA χB P∧Q PH ∩ QH
A⊂B χA χB = χA P≤Q PH ⊂ QH
k=1 Pk H
S∞ S∞
k=1 Ak χ∪k A ∨∞
k=1 Pk span{ }
k=1 Pk H
T∞ T∞
k=1 Ak χ∩k Ak ∧∞
k=1 Pk
PH ⊂ QH (1.24)
m
P = PQ (1.25)
m
P ≤ Q. (1.26)
2 Earlier authors, Schmidt and Hilbert, worked with infinite bases, and ∞ × ∞ matrices.
1.6 The Lattice Structure of Projections 51
and P really defines a selfadjoint projection. (We use “≤” to denote the lattice op-
eration on projection.) Note the convergence refers to the strong operator topology,
i.e., for all x ∈ H , there exists a vector, which we denote by Px, so that
52 1 Elementary Facts
The examples in section 1.4 using Gram-Schmidt process can now be formulated
in the lattice of projections.
Recall that for a linearly independent subset {uk } ⊂ H , the Gram-Schmidt pro-
cess yields an orthonormal set {vk } ⊂ H , with v1 := u1 / ku1 k, and
un − Pn un
vn+1 := , n = 1, 2, . . . ;
kun − Pn un k
∨Pn = sup Pn = I
∧Pn⊥ = inf Pn⊥ = 0.
and so
(P + Q)2 = P + Q (1.32)
m
PQ + QP = 0. (1.33)
Then,
(PQ)2 = P (QP) Q = −P(PQ)Q = −PQ
(1.34)
PQ = 0 or PQ = I. (1.35)
h f , χE f i ≤ h f , χF f i , ∀ f ∈ L2 (X) .
Lemma 1.13. Let P, Q ∈ Pro j (H ); then the following conditions are equivalent:
1. PQP ∈ Pro j (H );
2. PQ = QP.
For readers wishing to follow up sources, or to go in more depth with topics above,
we suggest: [Arv72, Ban93, BR79, Con90, DM85, DS88c, Lax02, RS75, RSN90,
Rud73, Rud87, AJS14, AJLM13, AJL13, AJ12].
Chapter 2
Unbounded Operators in Hilbert Space
— Max Born
55
56 2 Unbounded Operators in Hilbert Space
001
0ac
of matrices 0 0 b, (a, b, c) ∈ R3 . Let U be a strongly continuous unitary repre-
000
sentation of G acting on a Hilbert space H . Then for every X ∈ g, t 7−→ U etX
defines a strongly continuous one-parameter group; and hence its infinitesimal gen-
erator, denoted dU (X) is a skew-adjoint operator with dense domain in H .
In the example below, we apply this to H = L2 (R), and the unitary representa-
tion U of G is defined as follows: For f ∈ H = L2 (R), set
called the Schrödinger representation. Differentiate in the three directions in the Lie
algebra, we get
(dU (X1 ) f ) (x) = f 0 (x) = dx
d
f,
(dU (X2 ) f ) (x) = ix f (x) , and (2.4)
(dU (X ) f ) (x) = i f (x) .
3
dU (X1 ) = iP
Remark 2.1. It would appear that the function fλ (x) = eiλ x , λ fixed is an eigenfunc-
1 d
tion for P = i dx . But then, for all λ ,
P fλ = λ fλ
holds pointwise, “spectrum” depends on the ambient Hilbert space H , in this case
H = L2 (R); and fλ ∈
/ L2 (R). Nonetheless, if we allow an interval for the λ vari-
able, e.g., a < λ < b, with a and b being finite, then
ˆ b
eibx − eiax
Fa,b (x) = eiλ x dλ =
a ix
is in L2 (R); and hence P has continuous spectrum. The functions Fa,b (·) are exam-
ples of wave-packets in quantum mechanics.
Example 2.1. d/dx and Mx in QM, acting on L2 (R) with dense domain the Schwartz
space.
An alternative way to get a dense domain, a way that works for all representations,
is to use Gårding space, or C∞ vectors. Let u ∈ H and define
ˆ
uϕ := ϕ(g)Ug udg
G
≤ ϕε (g) u −Ug u dg
G
where the integration on G is w.r.t. Haar measure, and where we used the fact that
´
G ϕε = 1. Notice that we always assume the representations are norm continuous
58 2 Unbounded Operators in Hilbert Space
Probability Kernels. There are about 7 popular kernels in probability theory. Look
1 1
at any book on probability theory. One of them is the Cauchy kernel π 1+x2 . Notice
that not only uϕ is dense in H , their derivatives are also dense in H .
1. A = Ā.
2. G (A) = G (A).
3. dom (A) is a Hilbert space with respect to the graph inner product h·, ·iA .
4. If (an , Aan ) is a sequence in G (A) and (an , Aan ) → (a, b), then (a, b) ∈
G (A). In particular, b = Aa.
Let X be a vector space over C. Suppose there are two norms defined on X, such
that
k·k1 ≤ k·k2 . (2.5)
Let Xi be the completion of X with respect to k·ki , i = 1, 2. The ordering (2.5) implies
the identify map
ϕ : (X, k·k2 ) → (X, k·k1 )
Lemma 2.2. k·k1 and k·k2 are topologically equivalent if and only if
{xn } ⊂ X is Cauchy under k·k2
(hence Cauchy under k·k1 ) =⇒ kxn k2 → 0.
kxn k1 → 0
1
(kxk + kAxk)2 ≤ kxk2 + kAxk2 ≤ (kxk + kAxk)2 , ∀x ∈ dom (A) .
2
Theorem 2.2. An operator A is closable if and only if k·k and k·kA are topologi-
cally consistent. (When they do, the completion of dom (A) with respect to k·kA is
identified as a subspace of H .)
Proof. First, assume A is closable. Let {xn } be a sequence in dom (A). Suppose {xn }
is a Cauchy sequence with respect to k·kA , and kxn k → 0. We need to show {xn }
converges to 0 under the A-norm, i.e., kxn kA → 0. Since {(xn , Axn )} ⊂ G (A), and
A is closable, it follows that (xn , Axn ) → (0, 0) ∈ G (A). Therefore, kAxn k → 0, and
(see lemma 2.3)
kxn kA = kxn k + kAxn k → 0.
Conversely, assume k·k and k·kA are topologically consistent. Let {xn } ⊂ dom (A),
such that
(xn , Axn ) → (0, b) in H ⊕ H . (2.6)
Corollary 2.1. An operator A with dense domain is closable if and only if its adjoint
A∗ has dense domain.
We will focus on unbounded operators. In the sequel, we will consider densely de-
fined Hermitian (symmetric) operators. Such operators are necessarily closable.
The following result is usually applied to operators whose inverses are bounded.
Proposition 2.1. Let A be a bounded operator with domain dom (A), and act in H .
Then dom (A) is closed in k·kA if and only if it is closed in k·k. (That is, for bounded
operators, k·k and k·kA are topologically equivalent. )
60 2 Unbounded Operators in Hilbert Space
Corollary 2.2. If A is a closed operator in H and A−1 is bounded, then ran (A) is
closed in both k·k and k·kA−1 .
Proof. Note the G (A) is closed iff G A−1 is closed; and ran (A) = dom A−1 .
1. D (A) is dense in H .
2. (b, 0) ⊥ G (A) =⇒ b = 0.
3. If (b, −b∗ ) ⊥ G (A), the map b 7→ b∗ is well-defined.
for all a ∈ D (A). G (A)⊥ is the inverted graph of A∗ . The adjoints are only defined
for operators with dense domains in H .
D = f ∈ C1 f (0) = f (1) = 0 .
For unbounded operators, (AB) ∗ = B∗ A∗ does not hold in general. The situation is
better if one of them is bounded.
2.2 Characteristic Matrix 61
The next theorem follows directly from the definition of the adjoint operators.
• selfadjoint if A = A∗ .
• essentially selfadjoint if A = A∗ .
• normal if A∗ A = AA∗ .
• regular if D (A) is dense in H , and closed in k·kA .
Definition 2.2. Let A be a linear operator on a Hilbert space H . The resolvent R (A)
is defined as
n o
R (A) = λ ∈ C : (λ − A)−1 exists (the resolvent set)
Exercise 2.1. Let A be a linear operator in a Hilbert space H , and, for λi ∈ R (A),
i = 1, 2 consider two operators (λi − A)−1 . Show that
∑ Pik Pk j = Pi j (2.9)
k
If any of these conditions is satisfied, let A be the operator with G (A) = K , then
for all a, b ∈ H ,
" #" # " #
P11 P12 a P11 a + P12 b
= ∈ G (A) ;
P21 P22 b P21 a + P22 b
i.e.,
A : (P11 a + P12 b) 7→ P21 a + P22 b. (2.10)
In particular,
Theorem 2.6. Let A be an operator with characteristic matrix P = (Pi j ). The fol-
lowing are equivalent.
1. D (A) is dense in H .
2. " #
b
⊥ G (A) = 0 =⇒ b = 0.
0
" #
−b∗
3. If ∈ G (A)⊥ , the map A∗ : b 7→ b∗ is a well-defined operator.
b
4. " #" # " #
1 − P11 −P12 b b
= =⇒ b = 0.
−P21 1 − P22 0 0
5.
P11 b = 0, P21 b = 0 =⇒ b = 0.
that is,
A∗ : P21 a − (1 − P22 )b 7→ (1 − P11 )a − P12 b. (2.13)
In particular,
Proof. (1) ⇔ (2) ⇔ (3) is a restatement of Proposition 2.2. Note the projection
⊥
from H ⊕ H on G (A)⊥ = G (A) is
" #
1 − P11 −P12
1−P =
−P21 1 − P22
Theorem 2.7. Let A be a regular operator (i.e., densely defined, closed) with char-
acteristic matrix P = (Pi j ).
1. The matrix entries Pi j are given by
−1
That is, 1 + A∗ A is a Hermitian extension of P11 . By (2.8), P11 is selfadjoint and so
−1 −1
is P11 . Therefore, 1 + A∗ A = P11 , or
P11 = (1 + A∗ A)−1 .
By (2.11),
P21 = AP11 = A(1 + A∗ A)−1 .
This means 1 + AA∗ ⊃ (1 − P22 )−1 is a Hermitian extension of the selfadjoint oper-
ator (1 − P22 )−1 (note P22 is selfadjoint), hence 1 + AA∗ is selfadjoint, and
By (2.15),
P12 = A∗ (1 − P22 ) = A∗ (1 + AA∗ )−1 .
By (2.12),
P22 = AP12 = AA∗ (1 + AA∗ )−1
∗
P21 = P12 = (A∗ (1 − P22 ))∗ ⊃ (1 − P22 )A
gives (2.18).
2.2.1 Commutants
Proof. Suppose BA ⊂ AB, we check that BA ⊂ AB. The converse is trivial. For
(a, Aa) ∈ G (A), choose a sequence (an , Aan ) ∈ G (A) such that (an , Aan ) → (a, Aa).
By assumption, (Ban , ABan ) = (Ban , BAan ) ∈ G (A). Thus, (Ban , ABan ) → (Ba, BAa) ∈
G (A). That is, Ba ∈ D(A) and ABa = BAa.
Lemma 2.5. Let A be a closed operator with characteristic matrix P = (Pi j ). Let B
be a bounded operator, and " #
B0
QB := .
0B
Proof. Obvious.
66 2 Unbounded Operators in Hilbert Space
Remark 2.2. Let M be a von Neumann algebra. Let x ∈ M s.t. kxk ≤ 1 and x = x∗ .
√
Set y := x + i 1 − x2 . Then, y∗ y = yy∗ = x2 + 1 − x2 = 1, i.e., y is unitary. Also,
x = (y + y∗ ) /2.
Theorem 2.8. Let A be a closed operator with characteristic matrix P = (Pi j ). Let
M be a von Neumann algebra, and
" #
B0
QB := , B ∈ M0 .
0B
1. A is affiliated with M.
2. PQB = QB P, for all B ∈ M0 .
3. Pi j ∈ M.
4. If D(A) is dense, then A∗ is affiliated with M.
1. A∗ A is selfadjoint;
2. D(A∗ A) is a core of A, i.e.,
A D (A∗ A)
= A;
2.3 Normal Operators 67
R(1 + A∗ A) = D(P11 ) = H .
Proof. Suppose A is normal. Then for all a ∈ D (A∗ A) (= D (AA∗ )), we have
kAak2 = hAa, Aai = ha, A∗ Aai = ha, AA∗ ai = hAa, A∗ ai = kA∗ ak2 ;
i.e., kAak = kA∗ ak, for all a ∈ D (A∗ A). It follows that
D A D (A∗ A)
= D A∗ D (AA∗ )
.
Theorem 2.11 (M.S. Stone). Let A be a regular operator in a Hilbert space H . Let
P = (Pi j ) be the characteristic matrix of A. The following are equivalent.
1. A is normal.
2. Pi j are mutually commuting.
68 2 Unbounded Operators in Hilbert Space
Remark 2.3. For the equivalence of 1 and 2, we refer to the original paper of Stone.
The most interesting part is 1 ⇔ 3. The idea of characteristic matrix gives rise to an
elegant proof without reference to the spectral theorem.
Theorem 2.12.
√
1. |A| := A∗ A is the unique positive selfadjoint operator T satisfying D(T ) =
D(A), and kTak = kAak for all a ∈ D(A).
2. ker (|A|) = ker (A), R(|A|) = R(A∗ ).
√
Proof. Suppose T = A∗ A, i.e. T ∗ T = A∗ A. Let D := D(T ∗ T ) = D(A∗ A). By The-
orem 2.9, D is a core of both T and A. Moreover, kTak = kAak, for all a ∈ D.
We conclude from this norm identity that D(T ) = D(A) and kTak = kAak, for all
a ∈ D(A).
Conversely, suppose T has the desired properties. For all a ∈ D(A) = D(T ), and
b ∈ D(A∗ A),
hT b, Tai = hAb, Aai = hA∗ Ab, ai
This implies that T b ∈ D(T ∗ ) = D(T ), T 2 b = A∗ Ab, for all b ∈ D(A∗ A). That is,
T 2 is a selfadjoint extension of A∗ A. Since A∗ A is selfadjoint, T 2 = A∗ A.
The second part follows from Theorem 2.4.
A = V |A| . (2.19)
Equation (2.19) is called the polar decomposition of A. It is clear that such decom-
position is unique.
Taking adjoints in (2.19) yields A∗ = |A|V ∗ , so that
Restrict AA∗ to R(A), and restrict A∗ A restricted to R(A∗ ). Then the two restrictions
are unitarily equivalent. It follows that A∗ A, AA∗ have the same spectrum, aside from
possibly the point 0.
√
By (2.20), |A∗ | = V |A|V ∗ = VA∗ , where |A∗ | = AA∗ . Apply V ∗ on both sides
gives
A∗ = V ∗ |A∗ | . (2.21)
Theorem 2.13. A is affiliated with a von Neumann algebra M if and only if |A| is
affiliated with M and V ∈ M.
Proof. Let U be a unitary operator in M0 . The operator UAU ∗ has polar decompo-
sition
UAU ∗ = (UVU ∗ )(U |A|U ∗ ).
For readers wishing to follow up sources, or to go in more depth with topics above,
we suggest: [BR81, DS88c, Jor08, Kat95, KR97b, Sto51a, Sto51b, Wei03, Yos95].
Chapter 3
The Spectral Theorem
Most Functional Analysis books, when covering the Spectral Theorem, stress the
bounded case. Because of dictates from applications (especially quantum physics),
below we stress questions directly related to key-issues for unbounded linear op-
erators. These themes will be taken up again in chapters 9 and 10. In a number of
applications, some operator from physics may only be “formally selfadjoint” also
called Hermitian; and in such cases, one asks for selfadjoint extensions (if any),
chapter 9. Chapter 10 is a particular case in point, arising in the study of infinite
graphs.
3.1 An Overview
von Neumann’s spectral theorem states that an operator A acting in a Hilbert space
H is normal if and only if there exits a projection-valued measure on C so that
ˆ
A= zPA (dz) (3.1)
sp(A)
71
72 3 The Spectral Theorem
Remark 3.1. Let ϕ : R → R be measurable and let A = A∗ be given; then, for ev-
(A)
ery f ∈ H \ {0}, set dµ f (λ ) := kPA (dλ ) f k2 ∈ M+ (R) (the finite positive Borel
measures on R.) Then the transformation formula (3.3) takes the following equiva-
lent form:
(ϕ(A)) (A)
dµ f = dµ f ◦ ϕ −1 , i.e., (3.4)
(ϕ(A)) (A)
dµ f (4) = dµ f ◦ ϕ −1 (4) , ∀4 ∈ B (R) . (3.5)
f ∈ dom (ϕ (A)) ⇐⇒ ϕ ∈ L2 R, µ f ,
Quantum mechanics is stated using an abstract Hilbert space as the state space. In
practice, one has the freedom to choose exactly which Hilbert space to use for a par-
ticular problem. Physical measurements remain unchanged when choosing different
realizations of a Hilbert space. The concept needed here is unitary equivalence.
i.e., the observable P has the same expectation value. Since every selfadjoint oper-
ator is, by the spectral theorem, decomposed into selfadjoint projections, it follows
that the expectation value of any observable remains unchanged under unitary trans-
formations.
We will also consider family of selfadjoint operators. Heisenberg’s commutation
relation PQ − QP = −iI is an important example of two non-commuting selfadjoint
operators.
Example 3.2. Let Q and P be the position and momentum operators in quantum
mechanics. That is, Q = Mx = multiplication by x, and P = −id/dx both defined
on the Schwartz space S (R)–space of rapidly decreasing functions on R, which
is dense in the Hilbert space L2 (R). On S (R), the operators P and Q satisfy the
canonical commutation relation: PQ − QP = −iIL2 (R) .
Denote F the Fourier transform on L2 (R) as before. Specifically, setting
ˆ
1
(Fϕ) (x) = ϕb (x) = √ ϕ (ξ ) e−iξ x dξ , and
2π R
74 3 The Spectral Theorem
ˆ
1
(F ∗ ψ) (ξ ) = ψ ∨ (ξ ) = √ ψ (x) eiξ x dx, ξ ∈ R.
2π R
1 d
(F ∗ QFϕ) (ξ ) = F ∗ (xϕb (x)) = ϕ (ξ ) , ∀ϕ ∈ S.
i dξ
Therefore,
P = F ∗ QF (3.6)
Example 3.3. Eq. (3.6) says that P is diagonalized by Fourier transform in the fol-
lowing sense.
Let ψ be any Borel function on R, and set Mψ = multiplication by ψ (x) in
L2 (R), with
dom Mψ = f f , ψ f ∈ L2 (R)
(3.7)
ˆ ∞
= f 1 + |ψ (x)|2 | f (x)|2 dx < ∞ ;
−∞
ψ (P) := F ∗ ψ (Q) F.
E(4) = F ∗ Mχ4 F.
One checks directly that E (4)2 = E (4) = E (4)∗ , so E (4) is a selfadjoint pro-
jection. Indeed, E (·) is a convolution operator, where
ˆ b ∧
(E (4) f ) (x) = eiξ x fb(ξ ) dξ = f ∗ χ[a,b] (x) , ∀ f ∈ L2 (R) .
a
3.1 An Overview 75
Thus,
n o
E (4) L2 (R) = f ∈ L2 (R) supp fb ⊂ 4 ;
Example 3.4. Below, it helps to denote the Fourier transformed space (or frequency
space) by L2 (R).
b Fix any f ∈ L2 (R), 4 ∈ B (R), then
The last step above yields the projection-valued measure (PVM) version of the spec-
tral theorem for P, where we write
76 3 The Spectral Theorem
ˆ ∞
P= x dE (x) . (3.8)
−∞
1 d
Consequently, we get two versions of the spectral theorem for P = i dx S(R) :
This example illustrates the main ideas of the spectral theorem of a single selfadjoint
operator in Hilbert space. We will develop the general theory in this chapter, and
construct both versions of the spectral decomposition.
Example 3.5. Applying the Gram-Schmidt process to all polynomials against the
2 /2
measure dµ = e−x dx, one gets orthogonal polynomials Pn in L2 (µ). These are
called the Hermite polynomials, and the associated Hermite functions are given by
n
−x2 /2 −x2 d 2 /2
hn := e Pn = e ex .
dx
The Hermite functions (after normalization) forms an ONB in L2 (R), which trans-
forms P and Q to Heisenberg’s infinite matrices in (1.18)-(1.19).
1
H := (Q2 + P2 − 1).
2
or equivalently,
(P2 + Q2 )hn = (2n + 1)hn
Example 3.7 (Purely discrete spectrum v.s. purely continuous spectrum). The two
operators P2 + Q2 and P2 − Q2 acting in L2 (R); see Fig 3.1.
3.2 Multiplication Operator Version 77
In the following sections, we present some main ideas of the spectral theorem for
single normal operators acting in Hilbert space. Since every normal operator N can
be written as N = T1 + iT2 , where T1 and T2 are strongly commuting and selfadjoint,
the presentation will be focused on selfadjoint operators.
This version of the spectral theory states that every selfadjoint operator A is unitarily
equivalent to the operator of multiplication by a measurable function on some L2 -
space.
Theorem 3.1. Let A be a linear operator acting in the Hilbert space H , then A = A∗
iff there exists a unitary operator U : L2 (X, µ) → H such that
Mϕ = U ∗ AU; (3.9)
HO
A /H
O
U U
Mϕ
L2 (X, µ) / L2 (X, µ)
Exercise 3.2. Let Mt : L2 [0, 1] → L2 [0, 1], f (t) 7−→ t f (t). Show that Mt has no
eigenvalues in L2 [0, 1].
Before giving a proof of Theorem 3.1, we show below that one can go one step
further and get that A is unitarily equivalent to the operator of multiplication by
the independent variable on some L2 -space. This is done by a transformation of the
measure µ in (3.11).
µϕ := µ ◦ ϕ −1 (3.12)
Proof. For any simple function s = ∑ ci χEi , Ei ∈ TY , it follows from (3.13) that
3.2 Multiplication Operator Version 79
ˆ ˆ
s ◦ ϕdµ = ∑ ci χEi ◦ ϕdµ
X
ˆX
= ∑ ci χϕ −1 (Ei ) dµ
X
= ∑ ci µ ϕ −1 (Ei )
ˆ
= s d µ ◦ ϕ −1 .
Y
Remark 3.3. If ϕ is nasty, even if µ is a nice measure (say the Lebesgue measure),
the transformation measure µ ◦ ϕ −1 in (3.14) can still be nasty, e.g., it could even be
singular.
Lemma 3.2. In Theorem 3.1, assume A is bounded selfadjoint, so that ϕ ∈ L∞ (X, µ),
and real-valued. Let µϕ := µ ◦ ϕ −1 (eq. 3.12), supported on the essential range of
ϕ. Then the operator W : L2 R, µϕ −→ L2 (X, µ), by
(W f ) (x) = f (ϕ (x)) , ∀ f ∈ L2 µϕ
(3.15)
is isometric, and
W Mt = Mϕ W, (3.16)
where Mt : L2 Y, µ f −→ L2 Y, µ f , given by
ˆ ˆ
k f k2L2 (Y,µϕ ) = 2
| f | dµϕ = | f ◦ ϕ|2 dµ = kW f k2L2 (X,µ)
Y X
80 3 The Spectral Theorem
so W is isometric. Moreover,
Mϕ W f = ϕ (x) f (ϕ (x))
W Mt f = W (tg (t)) = ϕ (x) f (ϕ (x))
Mt = F ∗ AF ; (3.18)
H
A /H
? O O _
U U
Mϕ
F L2 (X, µ) / L2 (X, µ) F
O O
W W
L2 (R, µϕ )
Mt
/ L2 (R, µϕ )
Remark 3.4.
ψ(A) := π(ψ), ψ ∈ A.
∑ ck xk 7−→ ∑ ck Ak
is well-defined, and it extends to all bounded measurable functions.
π
4. To see that ψ −→ ψ(A) := FMψ F ∗ (eq. 3.19) is an algebra isomorphism,
one checks that
= ψ1 (A) ψ2 (A)
Let ϕ be as in Lemma 3.2. For the more general case when ϕ is not necessarily
invertible, so W as in (3.16) may not be unitary, we may still diagonalize A, i.e., get
that A is unitarily equivalent to multiplication by the independent variable in some
L2 -space; but now the L2 -space is vector-valued, and what is involved is a direct
integral representation. This argument is briefly sketched in the next section.
f 7−→ ϕ f . (3.21)
Note that ϕ is real-valued. Moreover, A is bounded iff Mϕ ∈ L∞ (X, µ). When the
Hilbert space H is separable, we may further assume that µ is finite, or take µ to
be a probability measure.
To further diagonalize Mϕ , in the case when ϕ is “nasty”, we will need the fol-
lowing tool from measure theory.
Definition 3.3. Let X be a locally compact and Hausdorff space, and µ a Borel
probability measure on X. Let ϕ : X → Y be a measurable function, and set ν :=
µ ◦ ϕ −1 . A disintegration of µ with respect to ϕ is a system of probability measures
µy : y ∈ Y on X, satisfying
where µy : y ∈ essential range of ϕ ⊂ R is the system of probability measures as
in Definition 3.3.
The RHS in (3.23) is the Hilbert space consisting of measurable cross-sections
S 2
f :R→ L (µy ), where f (y) ∈ L2 (µy ), ∀y, and with the inner product given by
ˆ
h f , giL2 (ν) := h f (y) , g (y)iL2 (µy ) dν (y) .
Y
i.e., taking disjoint union as u j runs through all the cyclic vectors. When
H is separable, we get H = ⊕ j∈N H j , and we may set µ := ∑∞j=1 2− j µ j .
Details below.
Lemma 3.3. There exists a family of cyclic vector {uα } s.t. H = ⊕α Huα , orthog-
onal sum of cyclic subspaces.
Lemma 3.4. Set K := [− kAk , kAk]. For each cyclic vector u, there exists a Borel
measure µu s.t. supp (µu ) ⊂ K; and Huα ' L2 (K, µu ).
84 3 The Spectral Theorem
Define W : Hu −→ L2 (K, µu ), by
= u, f¯(A) f (A)u H
∗
= hu, f (A) f (A)uiH
= h f (A)u, f (A)uiH
= k f (A)uk2H .
3.2 Multiplication Operator Version 85
WA f (A) = WA(a0 + a1 A + a2 A2 + · · · + an An )
= W (a0 A + a1 A2 + a2 A3 + · · · + an An+1 )
= a0t + a1t 2 + a2t 3 + · · · + ant n+1
= t f (t)
= Mt W f (A)
Lemma 3.6. There exists a locally compact Hausdorff space X and a Borel measure
µ, a unitary operator F : H −→ L2 (X, µ), such that
A = F ∗ Mϕ F
where ϕ ∈ L∞ (µ).
Proof. Recall that we get a family of states w j , with the corresponding measures
µ j , and Hilbert spaces H j = L2 (µ j ). Note that all the L2 -spaces are on K = sp(A).
So it’s the same underlying set, but with possibly different measures.
To get a single measure space with µ, Nelson [Nel69] suggested taking the dis-
joint union
[
X := K × { j}
j
and µ := the disjoint union of µ 0j s. The existence of µ follows from Riesz. Then we
get
F
H = ⊕H j −−→ L2 (X, µ).
Exercise 3.3. Prove that π ∈ Rep(L∞ (µ), L2 (µ)) is multiplicity free. Conclude that
each cyclic representation is multiplicity free (i.e., it is maximal abelian.)
Hint: Suppose B ∈ B L2 (µ) commutes with all Mϕ , ϕ ∈ L2 (µ). Define g = B1,
Bψ = Bψ1 = BMψ 1 = Mψ B1 = Mψ g = ψg = gψ = Mg ψ
thus B = Mg .
Finish the proof of Theorem 3.1 for the case when A is unbounded and selfadjoint.
CA := (A − i) (A + i)−1
is then unitary. See section 9.2. Apply corollary 3.4 to CA , and convert the result
back to A. See, e.g., [Nel69, Rud73].
A projection valued measure (PVM) P satisfies the usual axioms of measures (here
Borel measures) but with the main difference:
Remark 3.8. The notion of PVM extends the familiar notion of an ONB:
Let H be a separable Hilbert space, and suppose {uk }k∈N is a ONB in H ; then
for 4 ∈ B (R), set
3.3 Projection-Valued Measure (PVM) 87
Remark 3.9. P(E) = F χE F −1 defines a PVM. In fact all PVMs come from this way.
In this sense, the Mt version of the spectral theorem is better, since it implies the
PVM version. However, the PVM version facilitates some formulations in quantum
mechanics, so physicists usually prefer this version.
Remark 3.10. Suppose we start with the PVM version of the spectral theorem. How
to prove (ψ1 ψ2 )(A) = ψ1 (A)ψ2 (A)? i.e. how to check we do have an algebra iso-
morphism? Recall in the PVM version, ψ(A) is defined as the operator so that for
all ϕ ∈ H , we have ˆ
ψdµϕ = hϕ, ψ(A)ϕi .
As a standard approximation technique, once starts with simple or even step func-
tions. Once it is worked out for simple functions, the extension to any measurable
functions is straightforward. Hence let’s suppose (WLOG) that the functions are
simple.
Proof. Let
ψ1 = ∑ ψ1 (ti )χEi
ψ2 = ∑ ψ2 (t j )χE j
then
ˆ ˆ
ψ1 P(dx) ψ2 P(dx) = ∑ ψ1 (ti )ψ2 (t j )P(Ei )P(E j )
i, j
Remark 3.11. As we delve into Nelson’s lecture notes [Nel69], we notice that on
page 69, there is another unitary operator. By piecing these operators together is
precisely how we get the spectral theorem. This “piecing” is a vast generalization of
Fourier series.
3.3 Projection-Valued Measure (PVM) 89
and we have the Hilbert space L2 (µϕ ). Take Hϕ := span{ψ(A)ϕ : ψ ∈ L∞ (µϕ )}.
Then the map
ψ(A)ϕ 7→ ψ
Proof. We have
Remark 3.12. Hϕ is called the cyclic space generated by ϕ. Before we can construct
Hϕ , we must make sense of ψ(A)ϕ.
Since EAE is a Hermitian matrix on M, we may apply the spectral theorem for finite
dimensional space and get
EAE = ∑ λk Pλk
where λk0 s are eigenvalues associated with the projections Pλk . It follows that
p(A)u = p(∑ λk Pλk )u = ∑ p(λk )Pλk u
90 3 The Spectral Theorem
and
2
kp(A)uk2 = ∑ |p(λk )|2 Pλk u
2
≤ max |p(t)| ∑ Pλk u
= max |p(t)|
since
2
∑ Pλk u = kuk2 = 1.
where we may define the operator ψ(A) so that ψ(A)u is the limit of pn (A)u.
Proof. The unique part follows from a standard argument. We will only prove the
existence of P.
Let F : L2 (µ) → H be the unitary operator so that A = FMt F ∗ . Define
P (E) := F χE F ∗
= P (E1 ) P (E2 ) .
´
Thus A = tP(dt).
Remark 3.14. In fact, A is in the closed (under norm or strong topology) span of
{P(E) : E ∈ B}. This is equivalent to say that t = F −1 AF is in the closed span of
the set of characteristic functions, the latter is again a standard approximation in
measure theory. It suffices to approximate tχ[0,∞] .
The wonderful idea of Lebesgue is not to partition the domain, as was the case
in Riemann integral over Rn , but instead the range. Therefore integration over an
arbitrary set is made possible. Important examples include analysis on groups.
Proposition 3.1. Let f : [0, ∞] → R, f (x) = x, i.e. f = xχ[0,∞] . Then there exists a
sequence of step functions s1 ≤ s2 ≤ · · · ≤ f (x) such that limn→∞ sn (x) = f (x).
sn (x) ≡ i2−n ≤ x
sn (x) + 2−n ≡ (i + 1)2−n > x
sn (x) ≤ sn+1 (x).
Corollary 3.5. Let f (x) = xχ[0,M] (x). Then there exists a sequence of step functions
sn such that 0 ≤ s1 ≤ s2 ≤ · · · ≤ f (x) and sn → f uniformly, as n → ∞.
Then
lim |( f (x) − sn (x))h(x)|2 = 0
n→∞
and
|( f (x) − sn (x))h(x)|2 ≤ const|h(x)|2 .
94 3 The Spectral Theorem
or equivalently,
k( f − sn )hk2 → 0
Quantum States.
´
Exercise 3.6. Let H = L2 (R), and f ∈ L2 (R) given, with k f k2 = | f (x)|2 dx = 1.
Suppose f ∈ dom (P) ∩ dom (Q), where P and Q are the momentum and position
operators, respectively. Show that
1
v f (P) v f (Q) ≥ . (3.28)
4
1
σ f (P) σ f (Q) ≥
2
p p
where σ f (P) = v f (P), and σ f (Q) = v f (Q).
3.5 The Spectral theorem for compact operators 95
3.5.1 Preliminaries
The setting for the first of the Spectral Theorems we will consider is as following:
(Restricting assumptions will be relaxed in subsequent versions!)
Let H be a separable (typically infinite dimensional Hilbert space assumed
here!) and let A ∈ B (H ) \ {0} be compact and selfadjoint, i.e., A = A∗ , and A
is in the k·kUN -closure of F R (H ). See section 1.5.1.
Theorem 3.5. With A as above, there is an ONB {uk }k∈N , and a sequence {λk }k∈N
s.t. |λ1 | ≥ |λ2 | ≥ · · · ≥ |λk | ≥ |λk+1 | ≥ · · · , limk→∞ λk = 0, and
1. Auk = λk uk , k ∈ N,
2. A = ∑∞
k=1 λk |uk ih uk |,
3. spec (A) = {λk } ∪ {0},
dim v ∈ H Av = λk v < ∞, for all k ∈ N.
4.
The ONB {uk } contains an ONB (possibly infinite) subset, an ONB for the sub-
space
Ker (A) = v ∈ H Av = 0 .
1 3 k k
f (x, y) = ∑ i f i x + y, ik x + y .
4 k=0
then
96 3 The Spectral Theorem
3
4 f (x, y) = ∑ ik f ik x + y, ik x + y
k=0
hx, Axi ∈ R, ∀x ∈ H .
Proof. Suppose A is selfadjoint, i.e., hx, Ayi = hAx, yi, ∀x, y ∈ H . Setting x = y, we
get
hx, Axi = hAx, xi = hx, Axi =⇒ hx, Axi ∈ R, ∀x ∈ H .
Conversely, suppose hx, Axi ∈ R, ∀x ∈ H . Note that hx, Ayi and hAx, yi are both
sesquilinear forms defined on H , by theorem 3.6, they satisfy the corresponding
polarization identity. Thus,
and
4 hAx, yi = hA (x + y) , x + yi + i hA (ix + y) , ix + yi −
hA (−x + y) , −x + yi − i hA (−ix + y) , −ix + yi .
Since hx, Axi ∈ R, ∀x ∈ H , the RHSs of the above equations are equal, therefore,
hx, Ayi = hAx, yi, ∀x, y ∈ H , i.e., A is selfadjoint.
Theorem 3.7. Let H be a Hilbert space over C. Let f : H × H → C be a
sesquilinear form, and M := sup {| f (x, y)| : kxk = kyk = 1} < ∞. There exists a
unique bounded operator A in H , satisfying
Proof. Set f (x, y) = hx, Ayi, x, y ∈ H , in theorem 3.7. By the uniqueness part in the
theorem, A is recovered from f (·, ·).
Proof. Given A, f (x, y) := hx, Ayi is a sesquilinear form; and (3.32) follows imme-
diately from theorem 3.7.
Sine A is selfadjoint, the four inner products on RHS of the above equation are all
real-valued (3.7). Thus,
1
ℜ {hx, Ayi} = (hA (x + y) , x + yi − hA (−x + y) , −x + yi) .
4
where the second equality follows from the fact that eiθ hx, Ayi ∈ R. Then,
1
|hx, Ayi| = |hA (x + y) , x + yi − hA (−x + y) , −x + yi|
4
1
= |hA (x + y) , x + yi − hA (x − y) , x − yi|
4
3.5 The Spectral theorem for compact operators 99
1
≤ M kx + yk2 + kx − yk2
4
1
= kMk 2 kxk2 + 2 kyk2
4
= kMk , assuming kxk = kyk = 1;
therefore, by 3.9,
For readers wishing to follow up sources, or to go in more depth with topics above,
we suggest: [Con90, FL28, Kat95, Kre55, Lax02, LP89, Nel69, RS75, Rud73,
Sto51a, Sto51b, Yos95, HJL+ 13, DHL09, HKLW07].
Part III
Applications
Chapter 4
GNS and Representations
— Werner Heisenberg
Since states in quantum physics are vectors (of norm one) in Hilbert space; two
question arise: “Where does the Hilbert space come from?” And “What are the
algebras of operators from which the selfadjoint observables must be selected?” In
a general framework, we offer an answer below, it goes by the name “the Gelfand-
Naimark-Segal (GNS) theorem, ” which offers a direct correspondence between
states and cyclic representations.
A+ = b∗ b b ∈ A .
Example 4.1. Let X be a compact Hausdorff space, the algebra C (X) of all continu-
ous function on X is a C∗ -algebra under the sup-norm.
1Kadison et al. in 1950’s reduced the axioms of C∗ -algebra from about 6 down to just one (C∗ 3)
on the C∗ -norm.
103
104 4 GNS and Representations
(ON 1) s∗i s j = δi j 1;
(ON 2) ∑Ni=1 si s∗i = 1.
as follows:
←− (easy direction): Given π ∈ Rep (A, H ), u0 ∈ H , ku0 k = 1, set
∼
Hϕ = A/ b ∈ A ϕ (b∗ b) = 0
hΩ , π (a) Ω iHϕ , ∀a ∈ A.
Definition 4.3. Let ϕ ∈ S (A), we say it is a pure state iff ϕ ∈ extS (A) := the
extreme-points in S (A).
Facts 4.1
1. If A is a C∗ -algebra , then S (A) (⊂ A∗ ) is convex and weak *-compact.
2. Given ϕ ∈ S (A), and let πϕ ∈ Rep (A, H ) be the GNS-representation, see
(4.3); then
Example 4.4 (Pure states, cases where the full list is known!). ϕ ∈ S (A):
A extS (A)
Exercise 4.1. Using GNS, write down explicitly the irreducible representations of
the three C∗ -algebras in Table 4.1 corresponding to the listed pure states. (Hint: In
106 4 GNS and Representations
the case of C (X), the representations are one-dimensional, but in the other cases,
they are infinite-dimensional, i.e., dim Hπϕ = ∞.
Groups
Case 1. Groups contained in B (H ) where H is a fixed Hilbert space:
Exercise 4.2. Let G be a locally compact group with µ = a left-invariant Haar mea-
sure. Set (ρL (g) f ) (x) := f g−1 x , g, x ∈ G, f ∈ L2 (G, µ). Then show that ρL is a
A = ∑ Ag g (4.5)
g
where Ag ∈ C, and making A := C[G] into a ∗-algebra with the following two oper-
ations on finite sums as in (4.5): A × A −→ A, given by
! ! !
∑ Ag g ∑ Bh h := ∑ ∑ Ah Bk g (4.6)
g∈G h∈G g hk=g
4.2 The GNS Construction 107
and !∗
∑ Ag g := ∑ Ag g−1 . (4.7)
g∈G g∈G
Fact 4.1 There is a bijection between Repuni (G, H ) and Rep (C[G], H ) as fol-
lows: If π ∈ Repuni (G, H ), set πe ∈ Rep (C[G], H ):
!
πe ∑ Ag g := ∑ Ag π (g) (4.8)
g∈G g∈G
where the element ∑g∈G Ag g in (4.8) is a generic element in C[G], see (4.5), i.e., is
a finite sum with Ag ∈ C, for all g ∈ G.
Exercise 4.3. Fill in the proof details of the assertion in Fact 4.1.
Example 4.5. Let G be a group, considered as a countable discrete group (the count-
ability is not important). Set H = l 2 (G), and
Definition 4.6. Let G, and π ∈ Rep G, l 2 (G) be as in (4.9), and let πe ∈ Rep C[G], l 2 (G)
∗
(G) := the norm closure of πe (C[G]) ⊂ B l 2 (G) ;
Cred
The GNS construction is a general principle for getting representations from given
data in applications, especially in quantum mechanics. It was developed indepen-
dently by I. Gelfand, M. Naimark, and I. Segal around 1960s.
108 4 GNS and Representations
1. π(AB) = π(A)π(B)
2. π(A∗ ) = π(A)∗
Example 4.6. The multiplication version of the spectral theorem of a single self-
adjoint operator, say A acting on H , yields a representation of the algebra of
L∞ (sp (A)) (or C(sp (A))) as operators on H , where
The general question is given any ∗-algebra, where to get such a representation?
The answer is given by states. One gets representations from algebras vis states.
For abelian algebras, the states are Borel measures, so the measures come out as a
corollary of representations.
is a state. In fact, in the abelian case, all states are Borel probability measures.
We are aiming at a proof of the GNS theorem (Theorem 4.1), and a way to get more
general representations of ∗-algebras. Indeed, any representation is built up by the
cyclic representations (Def. 4.9), and each cyclic representation is in turn given by
a GNS construction.
Theorem 4.2. Give any representation π ∈ Rep(A, H ), there exists an index set J,
closed subspaces H j ⊂ H such that
1. Hi ⊥ H j , ∀i 6= j;
2. ∑⊕j∈J H j = H ; and
3. there exists cyclic vectors v j ∈ H j such that the restriction of π to H j is
cyclic.
Remark 4.3. The proof of 4.2 is very similar to the construction of orthonormal
basis (ONB); but here we get a family of mutually orthogonal subspaces. Of course,
if H j ’s are all one-dimensional, then it is a decomposition into ONB. Note that not
every representation is irreducible, but every representation can be decomposed into
direct sum of cyclic representations.
i.e., the cyclic subspace generated by v1 . If Hv1 6= H , then ∃v2 ∈ H \Hv1 , and the
cyclic subspace Hv2 , so that Hv1 and Hv2 are orthogonal. If Hv1 ⊕ Hv2 6= H , we
then build Hv3 and so on. Now use transfinite induction or Zorn’s lemma to show
the family of direct sum of mutually orthogonal cyclic subspaces is total. The final
step is exactly the same argument for the existence of an ONB of any Hilbert space.
Now we proceed to prove the GNS theorem (Theorem 4.1), which is restated below.
then
W : π1 (A) u1 7−→ π2 (A) u2 , A ∈ A (4.12)
110 4 GNS and Representations
π2 = W π1 . (4.13)
Remark 4.4. For the non-trivial direction, let ϕ be a given state on A, and we need
to construct a cyclic representation (π, Hϕ , uϕ ). Note that A is an algebra, and it
is also a complex vector space. Let us try to turn A into a Hilbert space and see
what conditions are needed. There is a homomorphism A → A which follows from
the associative law of A being an algebra, i.e., (AB)C = A(BC). To continue, A
should be equipped with an inner product. Using ϕ, we may set hA, Biϕ := ϕ (A∗ B),
∀A, B ∈ A. Then h·, ·iϕ is linear in the second variable, and conjugate linear in the
first variable. It also satisfies hA, Aiϕ = ϕ (A∗ A) ≥ 0. Therefore we take Hϕ :=
[A/{A : ϕ(A∗ A) = 0}]cl .
Thus ϕ is a state.
Conversely, fix a state ϕ on A. Set
( )
n
H0 := ∑ ci Ai ci ∈ C, n ∈ N
i=1
∑ ci Ai , ∑ di Bi ϕ
:= ∑∑ ci d j ϕ (A∗i B j ) .
2
∑ ci Ai ϕ
= ∑ ci Ai , ∑ ci Ai ϕ
= ∑∑ ci c j ϕ (A∗i A j ) ≥ 0. (4.14)
4.2 The GNS Construction 111
The RHS of (4.14) is positive since ϕ is a state. Recall that ϕ (A∗ A) ≥ 0, for all
A ∈ A, and this implies that for all n ∈ N, the matrix (ϕ (A∗i A j ))ni, j=1 is positive
definite, hence (4.14) holds.
Now, let Hϕ := completion of H0 under h·, ·iϕ modulo elements s s.t. kskϕ = 0.
See Lemma 4.1 below. Hϕ is the desired cyclic space, consisting of equivalence
classes [A], ∀A ∈ A. Next, let uϕ = [1A ] = equivalence class of the identity element,
and set
π (A) := [A] = [A1A ] = [A][1A ];
using the fact that ϕ (C∗ ) = ϕ (C)∗ , ∀C ∈ A. The lemma follows from the estimate
(4.15).
112 4 GNS and Representations
Example 4.8. Let A = C[0, 1]. Set ϕ : f 7→ f (0), so that ϕ ( f ∗ f ) = | f (0)|2 ≥ 0. Then,
and C [0, 1] / ker ϕ is one dimensional. The reason is that if f ∈ C[0, 1] such that
f (0) 6= 0, then we have f (x) ∼ f (0) since f (x)− f (0) ∈ ker ϕ, where f (0) represents
the constant function f (0) over [0, 1]. This shows that ϕ is a pure state, since the
representation has to be irreducible.
Exercise 4.7. Fill in the details in the above proof of the GNS theorem.
Using GNS construction we get the following structure theorem for abstract C∗ -
algebras. As a result, all C∗ -algebras are sub-algebras of B (H ) for some Hilbert
space H .
Theorem 4.4 (Gelfand-Naimark). Every C∗ -algebra (abelian or non-abelian) is
isometrically isomorphic to a norm-closed sub-algebra of B(H ), for some Hilbert
space H .
Proof. Let A be any C∗ -algebra, no Hilbert space H is given from outside. Let
S(A) be the states on A, which is a compact convex subset of the dual space A∗ .
Here, compactness refers to the weak ∗-topology.
We use Hahn-Banach theorem to show that there are plenty of states. Specifically,
∀a ∈ A, ∃ϕ ∈ A∗ such that ϕ(a) > 0. It is done first on the 1-dimensional subspace
tA 7→ t ∈ R,
These are the continuous linear functionals. The Hahn-Banach theorem implies that
for all v ∈ V , kvk 6= 0, there exists lv ∈ V ∗ such that l(v) = kvk. Recall the con-
struction is to define lv on one vector, then use transfinite induction to extend it to
V . Notice that V ∗ is always complete, even if V is an incomplete normed space. In
other words, V ∗ is always a Banach space.
Now V is embedded into V ∗∗ (as we always do this) via the mapping
V 3 v 7→ ψ (v) ∈ V ∗∗ , where
ψ (v) (l) := l (v) , ∀l ∈ V ∗ .
Example 4.9. Let X be a compact Hausdorff space. The algebra C(X) of all con-
tinuous functions on X with the sup norm, i.e., k f k∞ := supx∈X | f (x)|, is a Banach
space.
Let B be a Banach space and denote by B∗ its dual space. B∗ is a Banach space as
well, where the norm is defined by
Proof. This is proved by showing B∗1 is a closed subspace in ∏kxk=1 C, where the
latter is given its product topology, and is compact and Hausdorff.
As an application, we have
114 4 GNS and Representations
h 7→ hh, ·i ∈ H ∗
This can also be seen by noting that H is unitarily equivalent to l 2 (A), with some
index set A, and l 2 (A) is reflexive.
The set of all bounded operators B(H ) on H is a Banach space. We ask two
questions:
The first question is extremely difficult and we will discuss that later. For the present
section, we show that
B(H ) = T1 (H )∗
In general, we want to get rid of the assumption that ρ ≥ 0. This is done using
the polar decomposition, which we will consider in section 2.4 even for unbounded
operators. It is much easier for bounded operators: If A ∈ B (H ), A∗ A is positive,
√
selfadjoint, and so by the spectral theorem, we may take |A| := A∗ A. Then, one
checks that
D√ √ E
kAxk2 = hAx, Axi = hx, A∗ Axi = A∗ Ax, A∗ Ax = k|A| xk2 ,
thus
kAk = k|A|k (4.17)
and there is a partial isometry V : range (|A|) → range (A), and the following polar
decomposition holds:
A = V |A| (4.18)
We will come back to this point in section 2.4 when we consider unbounded opera-
tors.
A = ∑ λn | fn ih en | (4.19)
n
Proof. Using the polar decomposition A = V |A|, we may first diagonalize |A| with
respect to some ONB {en } as
A = V |A| = ∑ λn |Ven ih en | = ∑ λn | fn ih en |
n n
where fn := Ven .
With the above discussion, we may work, instead, with compact operators A :
H → H so that A is a trace class operator if |A| (positive, selfadjoint) satisfies
condition (4.20).
Note the RHS in (4.21) is independent of the choice of the ONB. For if { fn } is
another ONB in H , using the Parseval identity repeatedly, we have
∑ h fn , A fn i = ∑ ∑ h fn , em i hem , A fn i
n m
= ∑ ∑ h fn , em i hA∗ em , fn i
m n
= ∑ hA∗ em , em i
m
= ∑ hem , Aem i .
m
Proof. By Corollary 4.3, there exists ONBs {en } and { fn }, and A has a decomposi-
tion as in (4.19). Then,
We have used the fact that trace (A) is independent of the choice of an ONBs.
Proof. Let l ∈ T1∗ . How to get an operator A? It is supposed to be the case such that
l(ρ) = trace(ρA)
118 4 GNS and Representations
for all ρ ∈ T1 . How to pull an operator A out of the hat? The idea also goes back
to Dirac. It is in fact not difficult to find A. Since A is determined by its matrix, it
suffices to fine h f , A f i which are the entries in the matrix of A.
l(| f1 ih f2 |) = h f2 , A f1 i .
Now we check that l(ρ) = trace(ρA). By Corollary 4.3, any ρ ∈ T1 can be written
as ρ = ∑n λn | fn ih en |, where {en } and { fn } are some ONBs in H . Then,
trace (ρA) = trace | f
∑ n n n
λ ih e | A
n
= trace ∑ λn |A fn ih en |
n
= ∑ ∑ λn hum , A fn i hen , um i
m n
= ∑ λn ∑ hum , A fn i hen , um i
n m
= ∑ λn hen , A fn i (= l (ρ))
n
where {un } is an ONB in H , and the last step follows from Parseval’s identity.
Remark 4.5. If B is the dual of a Banach space, then we say that B has a predual.
For example l ∞ = (l 1 )∗ , hence l 1 is the predual of l ∞ .
Another example: Let H 1 be hardy space of analytic functions on the disk
[Rud87]. (H 1 )∗ = BMO, where BMO refers to bounded mean oscillation. It was
developed by Charles Fefferman in 1974 who won the fields medal for this theory.
See [Fef71]. (Getting hands on a specific dual space is often a big thing.)
Definition 4.12. Let f be a locally integrable function on Rn , and let Q run through
all n-cubes ⊂ Rn . Set ˆ
1
fQ = f (y) dy.
|Q| Q
In this case the LHS in (4.22) is the BMO-norm of f . Moreover, BMO is a Banach
space.
Below we consider some cases of building new Hilbert spaces from given ones.
Only sample cases are fleshed out; and they will be needed in the sequel.
(a) Let Hi , i = 1, 2 be two given Hilbert spaces, then the direct “orthogonal” sum
H = H1 ⊕ H2 is as follows:
(b) Given an indexed family of Hilbert spaces {Hα }α∈A where A is a set; then set
H := ⊕A Hα to be
2
⊕
∑ hα = ∑ khα k2Hα < ∞; (4.24)
α∈A H α∈A
H S (H ) := T ∈ B (H ) T ∗ T is trace class
(4.25)
and set
kT k2H S := trace (T ∗ T ) ; (4.26)
similarly if S, T ∈ H S (H ), set
Case 4. Tensor-product
Let H1 and H2 be two Hilbert spaces, and consider finite-rank operators (rank-1
in this case):
h1 ⊗ h2 ←→ |h1 ih h2 | . (4.29)
Set
kh1 ⊗ h2 k2 := trace (T ∗ T ) = kh1 k2H1 kh2 k2H2 . (4.30)
For the Hilbert space H1 ⊗ H2 we take the H S-completion of the finite rank opera-
tors spanned by the set in (4.28). The tensor product construction fits with composite
system in quantum mechanics.
On the subspace
R (T ) = T h1 h1 ∈ H1
(4.32)
and Naimark, every abelian C∗ -algebra containing the identity element is isomor-
phic to the algebra C (X) of continuous functions on some compact Hausdorff space
X, which is unique up to homeomorphism. The classification of all the represen-
tations abelian C∗ -algebras, therefore, amounts to that of C(X). This problem can
be understood using the idea of σ -measures (square densities). It also leads to the
multiplicity theory of selfadjoint operators. The best treatment on this subject can
be found in [Nel69].
Here we discuss Gelfand’s theory on abelian C∗ -algebras. Throughout, we as-
sume all the algebras contain unit element.
ϕ : A → A/M → C, a 7→ a/M 7→ ta
Lemma 4.4. Let A be an abelian Banach algebra. If a ∈ A, and kak < 1, then 1A −a
is invertible.
Proof. Let a ∈ A, a 6= 0. Suppose λ := ϕ (a) s.t. |λ | > kak. Then ka/λ k < 1 and
so 1A − a/λ is invertible by Lemma 4.4. Since ϕ is a homomorphism, it must map
invertible element to invertible element, hence ϕ (1A − a/λ ) 6= 0, i.e., ϕ (a) 6= λ ,
which is a contradiction.
Let X be the set of all maximal ideals, identified with all homomorphisms in A∗1 ,
where A∗1 is the unit ball in A∗ . Since A∗1 is compact (see Banach-Alaoglu, Theorem
122 4 GNS and Representations
4.6), and X is closed in it, therefore X is also compact. Here, compactness refers to
the weak ∗-topology.
(ab)n = ∑ ak bn−k
k
a∗n = a−n
kak = ∑ |an |
n
1A = δ0 (Dirac mass at 0)
To identity X in practice, we always start with a guess, and usually it turns out to be
correct. Since Fourier transform converts convolution to multiplication,
ϕz
l 2 (Z) 3 a −−−−−→ ∑ an zn
ϕz (ab) = ∑(ab)n zn
= ∑ ak bn−k zn
n,k
= ∑ ak zk ∑ bn−k zn−k
k n
! !
= ∑ ak zk ∑ bk zk .
k k
4.6 States and Representation 123
Thus {z : |z| = 1} is a subspace in the Gelfand space X. Note that we cannot use
|z| < 1 since we are dealing with two-sided l 1 sequence. (If the sequences were
truncated, so that an = 0 for n < 0 then we allow |z| < 1. )
ϕz is contractive: |ϕz (a)| = |∑ an zn | ≤ ∑n |an | = kak.
Exercise 4.9. Prove that every homomorphism of l 2 (Z) is obtained as ϕz for some
|z| = 1. Hence X = {z : |z| = 1}.
Example 4.12. l ∞ (Z), with kak = supn |an |. The Gelfand space in this case is X =
β Z, the Stone-Cech compactification of Z, which are the ultrafilters on Z. β Z is
much bigger then p-adic numbers. Pure states on diagonal operators correspond to
β Z. See Chapter 8 for details.
Example 4.14. Let A be a ∗-algebra. Given two states s1 and s2 , by the GNS con-
struction, we get cyclic vectors ξi , and representations πi : A → B(Hi ), so that
si (A) = hξi , πi (A)ξi i, i = 1, 2. Suppose there is a unitary operator W : H1 → H2 ,
such that for all A ∈ A,
π1 (A) = W ∗ π2 (A)W.
Then
i.e., s2 (A) = s1 (A). Therefore the same state s = s1 = s2 has two distinct (unitarily
equivalent) representations.
Remark 4.6. A special case of states are measures when the algebra is abelian. Re-
call that all abelian C∗ -algebras with identity are C (X), where X is the correspond-
ing Gelfand space. Two representations are mutually singular π1 ⊥ π2 , if and only
if the two measures are mutually singular, µ1 ⊥ µ2 .
1. π is irreducible.
2. The commutant (π(A))0 is one-dimensional, i.e., (π(A))0 = cIA , c ∈ C.
4.6 States and Representation 125
Proof. Suppose (π(A))0 has more than one dimension. Let X ∈ (π(A))0 , then by tak-
ing adjoint, X ∗ ∈ (π(A))0 . X +X ∗ is selfadjoint, and X +X ∗ 6= cI since by hypothesis
(π(A))0 has more than one dimension. Therefore X + X ∗ has a non trivial spectral
/ {0, I}. Let H1 = P(E)H and H2 = (I − P(E))H .
projection P(E), i.e., P(E) ∈
H1 and H2 are both nonzero proper subspaces of H . Since P(E) commutes with
π(A), for all A ∈ A, it follows that H1 and H2 are both invariant under π.
Conversely, suppose (π(A))0 is one-dimensional. If π is not irreducible, i.e., π =
π1 ⊕ π2 , then for
" # " #
IH1 0 0 0
PH1 = , PH2 = 1 − PH1 =
0 0 0 IH2
we have
PHi π(A) = π(A)PHi , i = 1, 2
Corollary 4.6. π is irreducible if and only if the only projections in (π(A))0 are 0
or I.
Thus to test invariant subspaces, one only needs to look at projections in the com-
mutant.
If instead of taking the norm closure, but using the strong operator topology, ones
gets a von Neumann algebra. von Neumann showed that the weak closure of A is
equal to A00
If " #
X Y
∈ (π(A))0
UV
then
" #" # " #
X Y π1 (A) 0 Xπ1 (A) Y π2 (A)
=
UV 0 π2 (A) Uπ1 (A) V π2 (A)
" #" # " #
π1 (A) 0 X Y π1 (A)X π1 (A)Y
= .
0 π2 (A) UV π2 (A)U π2 (A)V
Hence
Therefore,
π1 (A)
/ H1
C H1 [
Y U U Y
π2 (A)
H2 / H2
is obtained by using a more general result, which relates t and selfadjoint operators
in the commutant (π(A))0 .
We now turn to characterize the relation between state and its GNS representa-
tion, i.e., specialize to the GNS representation. Given a ∗ algebra A, the states S(A)
forms a compact convex subset in the unit ball of the dual A∗ .
Let A+ be the set of positive elements in A. Given s ∈ S(A), let t be a positive
linear functional. By t ≤ s, we mean t(A) ≤ s(A) for all A ∈ A+ . We look for relation
between t and the commutant (π(A))0 .
t(·) = hΩ , π(·)AΩ i
t(a) = hΩ , π(a)AΩ i
= Ω , π(b2 )AΩ
= hΩ , π(b)∗ π(b)AΩ i
= hπ(b)Ω , Aπ(b)Ω i
≤ hπ(b)Ω , π(b)Ω i
= hΩ , π(a)Ω i
= s(a).
t(a) = hπ(b)Ω , ηi .
i.e., π(b)Ω 7→ t(b2 ) is a bounded quadratic form. Therefore, there exists a unique
A ≥ 0 such that
t(b2 ) = hπ(b)Ω , Aπ(b)Ω i .
1. t ≤ s ⇒ t = λ s for some λ ≥ 0.
2. π is irreducible.
3. s is a pure state.
t(a) = hΩ , π (a) AΩ i , ∀a ∈ A.
s(a) = hΩ , π (a) Ω i
= hΩ1 ⊕ Ω2 , π (a) Ω1 ⊕ Ω2 i
= hΩ1 , π (a) Ω1 i + hΩ2 , π (a) Ω2 i
2 Ω1 Ω1 2 Ω2 Ω2
= kΩ1 k , π (a) + kΩ2 k , π (a)
kΩ1 k kΩ1 k kΩ2 k kΩ2 k
2 Ω 1 Ω 1 2
Ω
2 Ω2
= kΩ1 k , π (a) + 1 − kΩ1 k , π (a)
kΩ1 k kΩ1 k kΩ2 k kΩ2 k
= λ s1 (a) + (1 − λ ) s2 (a) .
More general states in physics come from the mixture of particle states, which cor-
respond to composite system. These are called normal states in mathematics.
Let ρ ∈ T1 (H ) = trace class operator, s.t. ρ > 0 and tr(ρ) = 1. Define state
sρ (A) := tr(Aρ) , A ∈ B (H ). Since ρ is compact, by spectral theorem of compact
operators,
ρ = ∑ λk Pk
k
such that λ1 > λ2 > · · · → 0; ∑ λk = 1 and Pk = |ξk ih ξk |, i.e., the rank-1 projections.
(See section 4.3.) We have
sρ = ∑ λk sξk = ∑ λk |ξk ih ξk |
k k
Remark 4.9. Notice that tr(|ξ ihη|) = hη, ξ i. In fact, take any ONB {en } in H , then
where the last step follows from Parseval identity. (If we drop the condition ρ ≥ 0
then we get the duality (T1 H )∗ = B(H ). See Theorem 4.7.)
4.6 States and Representation 131
• states - unit vectors ξ ∈ H . These are all the pure states on B(H ).
• observable - selfadjoint operators A = A∗
• measurement - spectrum
The spectral theorem was developed by J. von Neumann and later improved by
Dirac and others. A selfadjoint operator A corresponds to a quantum observable,
and result of a quantum measurement can be represented by the spectrum of A.
ˆ z }|
ξλ
{
ξ = E(dλ )ξ
so that ˆ
2
kξ k = kE(dλ )ξ k2 = 1
Definition 4.16. A vector space is locally convex if it has a topology which makes
the vector space operators continuous, and if the neighborhoods {x + Nbh0 } have a
basis consisting of convex sets.
The context for Krein-Milman is locally convex topological spaces. It is in all func-
tional analysis books. Choquet’s theorem however comes later, and it’s not found
in most books. A good reference is the book by R. Phelps [Phe01]. The proof of
Choquet’s theorem is not specially illuminating. It uses standard integration theory.
K = conv(E(K)).
Note: The dual of a normed vector space is always a Banach space, so the theo-
rem applies. The convex hull in an infinite dimensional space is not always
closed, so close it. A good reference to locally convex topological space is
the lovely book by F. Treves [Tre67].
1 2
Example 4.15. Let (X, M, µ) be a measure space, where X is compact and Haus-
dorff. The set of all probability measures P(X) is a convex set. To see this, let
µ1 , µ2 ∈ P(X) and 0 ≤ t ≤ 1, then tµ1 + (1 − t)µ2 is a measure on X, moreover
(tµ1 + (1 − t)µ2 )(X) = t + 1 − t = 1, hence tµ1 + (1 − t)µ2 ∈ P(X). Usually we
don’t want all probability measures, but a closed subset.
Example 4.16. We compute extreme points in the previous example. K = P(X) is
compact convex in C(X)∗ , which is identified as the set of all measures due to Riesz.
134 4 GNS and Representations
C(X)∗ is a Banach space hence is always convex. The importance of being the dual
of some Banach space is that the unit ball is always weak ∗ compact (Banach-
Alaoglu, Theorem 4.6). Note the weak ∗ topology is just the cylinder/product topol-
ogy. The unit ball B∗1 sits inside the infinite product space (compact, Hausdorff)
∏v∈B,kvk=1 D1 , where D1 = {z ∈ C : |z| = 1}. The weak ∗ topology on B∗1 is just the
restriction of the product topology onto B∗1 .
Claim: E(K) = {δx : x ∈ X}, where δx is the Dirac measure supported at x ∈ X. By
´
Riesz, to know the measure is to know the linear functional. f dδx = f (x). Hence
we get a family of measures indexed by X. If X = [0, 1], we get a continuous family
of measures. To see these really are extreme points, we do the GNS construction on
the algebra A = C(X), with the state µ ∈ P(X). The Hilbert space so constructed
is simply L2 (µ). It’s clear that L2 (δx ) is 1-dimensional, hence the representation is
irreducible. We conclude that δx is a pure state, for all x ∈ X.
There is a bijection between state ϕ and Borel measure µ := µϕ ,
ˆ
ϕ(a) = a dµϕ .
X
√
since µ ∈ P (X) is a probability measure. Also, if f ≥ 0 then f = g2 , with g := f;
and ˆ
ϕ( f ) = g2 dµ ≥ 0.
Example 4.17. Let A = B(H ), and S (A) = states of A. For each ξ ∈ H , the map
A 7→ wξ (A) := hξ , Aξ i is a state, called vector state.
Claim: E(S) = vector states.
To show this, suppose W is a subspace of H such that 0 W H , and suppose
W is invariant under the action of B(H ). Then ∃h ∈ H , h ⊥ W . Choose ξ ∈ W . The
wonderful rank-1 operator (due to Dirac) T : ξ 7→ h given by T := |h ih ξ |, shows that
h ∈ W (since TW ⊂ W by assumption.) Hence h ⊥ h and h = 0. Therefore W = H .
We say B(H ) acts transitively on H .
4.7 Krein-Milman, Choquet, Decomposition of States 135
E(Pinv (Ω )) = [0, 1]
Intuitively, it says that the whole space X can’t be divided non-trivially into parts
where µ is invariant. The set X will be mixed up by the transformation σ .
Exercise 4.11 (Irrational rotation). Let θ > 0 be a fixed irrational number, and set
1.0 1.0
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0.2 0.4 0.6 0.8 1.0 0.2 0.4 0.6 0.8 1.0
√ √
(a) (x, 2x) mod 1, 1 ≤ x ≤ 5 (b) (x, 2x) mod 1, 0 ≤ x ≤ 15
dw
Then wK is a state, and wK w, i.e., w(A) = 0 ⇒ wK (A) = 0. We say that K = dwK
is a noncommutative Radon-Nikodym derivative.
Check:
wK (1) = 1
√ √
∗ w( KA∗ A K)
wK (A A) =
w(K)
√ √
w((A K)∗ (A K))
= ≥0
w(K)
The converse holds too [Sak71] and is called the noncommutative Radon-Nikodym
theorem.
4.7 Krein-Milman, Choquet, Decomposition of States 137
i.e. ˆ ˆ ˆ
f dµ = dFx ( f )dx = f (x)dx.
Equivalently, ˆ
dµ = δx dx
i.e. ˆ ˆ ˆ
f dµ = δx ( f )dx = f (x)dx.
i.e. ˆ
P() = P(·|start at x)dx.
Poisson integration.
138 4 GNS and Representations
(a ∗ b)n = ∑ ak bn−k
k
∗
(a )n = a−n
1A = δ0
F
a −−−−→ F (z) := ∑ an zn
Gelfand n
Homomorphism:
F
(l 1 , ∗) −
→ C(T1 )
(an ) 7→ F(z).
If we want to write F(z) as power series, then we need to drop an for n < 0.
Then F(z) extends to an analytic function over the unit disk. The sequence
space
{a0 , a1 , . . .}
The algebra L1 has no identity, but we may always insert one by adding
δ0 . So δ0 is the homomorphism f 7−→ f (0); and L1 (R) ∪ {δ0 } is again a
Banach ∗-algebra.
∗ g = fˆĝ.
where fd
vuv−1 = eiϕ u
vu = eiϕ uv
[p, q] = −iI
Exercise 4.12. Show that p, q and not be represented by bounded operators. Hint:
take the trace.
V1 (en ) = e2n
140 4 GNS and Representations
V2 (en ) = e2n+1
V1V1∗ +V2V2∗ = I
Vi∗Vi = I
ViVi∗ = Pi
The main question here is how to break up a representation into smaller ones. The
smallest representations are the irreducible ones. The next would be multiplicity
free representation.
Let A be an algebra.
Smallest representation:
where B ∈ M2 (C).
A = ∑ λn Pn
where λn0 s are the eigenvalues of A, such that λ1 ≥ λ2 ≥ · · · λn → 0. and Pn0 s are
self-adjoin projections onto the corresponding finite dimensional eigenspace of λn .
P is a PVM supported on {1, 2, 3, . . .}, so that Pn = P({n}).
Question: what does A look like if it is represented as the operator of multiplica-
tion by the independent variable?
We may arrange the eigenvalues of A such that
s s s
z }|1 { z }|2 { z }|n {
λ1 = · · · = λ1 > λ2 = · · · = λ2 > · · · > λn = · · · = λn > · · · → 0.
We say that λi has multiplicity si . The dimension of the eigen space of λi is si , and
dimH = ∑ si .
Notice that χEk is a rank s1 projection. M f is compact if and only if it is of the given
form.
M f χ{xi } = λ χ{xi } , i = 1, 2.
and
L2 (µ f ) = L2 (µ({x1 })δλ ) ⊕ L2 (µ({x2 })δλ ) ⊕ L2 (cont. sp δ0 ).
Define U : L2 (µ) → L2 (µ f ) by
(Ug) = g ◦ f −1 .
Mf
L2 (X, µ) / L2 (X, µ)
U U
L2 (R, µ f )
Mt
/ L2 (R, µ f )
To see U diagonalizes M f ,
Thus
Mt U = UM f .
Example 4.30. 2-d, λ I, {λ I}0 = M2 (C) which is not abelian. Hence mult(λ ) = 2.
(ϕ ⊗ I)(I ⊗ B) = ϕ ⊗ B
(I ⊗ B)(ϕ ⊗ I) = ϕ ⊗ B.
For readers wishing to follow up sources, or to go in more depth with topics above,
we suggest: [Arv76, BR79, KR97a, Mac85, Seg50].
Chapter 5
Completely Positive Maps
In the last chapter we studied two question from the use of algebras of operators
in quantum physics: “Where does the Hilbert space come from?” And “What are
the algebras of operators from which the selfadjoint observables must be selected?”
An answer is given in “the Gelfand-Naimark-Segal (GNS) theorem;” a direct cor-
respondence between states and cyclic representations. But states are scalar valued
positive definite functions on ∗-algebras. For a host of applications, one must instead
consider operator valued “states.” For this a different notion of positivity is needed,
“complete positivity.”
The GNS construction gives a bijection between states and cyclic representa-
tions. An extension to the GNS construction is Stinespring’s completely positive
map. It appeared in an early paper by Stinespring in 1956 [Sti55]. Arveson in 1970’s
reformatted Stinespring’s result using tensor product [Arv72]. He realized that com-
pletely positive maps are the key in multivariable operator theory, and in noncom-
mutative dynamics.
5.1 Motivation
w(A) = hΩ , π(A)Ω i
145
146 5 Completely Positive Maps
K = span{π(A)Ω : A ∈ A}.
∑ Ai j ⊗ ei j .
i, j
Remark 5.1. The algebra B(Cn ) of all bounded operators on Cn is generated by the
rank-one operators, i.e.,
IMn (ei j ) = ei ih e j . (5.4)
Hence the ei j on the LHS of (5.3) is seen as an element in the algebra Mn (C),
i.e., n × n complex matrices; while on the RHS of (5.3), ei j is treated as the rank one
operator ei ih e j ∈ B(Cn ). Using Dirac’s notation, when we look at ei j as operators,
we may write
|e i j=k
i
ei, j (ek ) = |ei ih e j | |ek i =
0 j 6= k
|e ihe | j=k
i l
ei, j ekl = |ei ihe j | |ek ihel | =
0 j 6= k
= ∑ vl , ϕ (Ai j ) v j hel , ei i
i, j,l
= ∑ vi , ϕ (Ai j ) v j ≥ 0. (5.5)
i, j
we have
ϕ(A11 ) ϕ(A12 ) · · · ϕ(A1n ) v1
ϕ(A21 ) ϕ(A22 ) · · · ϕ(A2n ) v2
i
h
v1 v2 · · · vn . .. .. .. . ≥ 0. (5.6)
.. . . . ..
ϕ(An1 ) ϕ(An2 ) · · · ϕ(Ann ) vn
ϕ(A) = hΩ , π (A) Ω iK
where
Ω = π (1A ) ∈ K
K = span{π(A)Ω : A ∈ A}.
V
C 3 t −→ tΩ ∈ CΩ (5.7)
VV ∗ : K → CΩ (5.8)
5.2 CP v.s. GNS 149
hξ ,V tiK = hV ∗ ξ ,tiC = tV ∗ ξ .
By setting t = 1, we get
V ∗ ξ = hξ ,V 1iK = hξ , Ω iK = hΩ , ξ iK ⇐⇒ V ∗ = hΩ , ·iK .
Therefore,
V ∗V t = V ∗ (tΩ ) = hΩ ,tΩ iK = t, ∀t ∈ C ⇐⇒ V ∗V = IC
VV ∗ ξ = V (hΩ , ξ iK ) = hΩ , ξ iK Ω , ∀ξ ∈ K ⇐⇒ VV ∗ = |Ω ih Ω | .
It follows that
ϕ (A) = hΩ , π (A) Ω iK
= hV 1, π (A)V 1iK
= h1,V ∗ π (A)V 1iC
= V ∗ π (A)V, ∀A ∈ A
In other words, Ω 7−→ π(A)Ω sends the unit vector Ω from the 1-dimensional
subspace CΩ to the vector π(A)Ω ∈ K , and hΩ , π (A) Ω iK cuts off the resulting
vector π(A)Ω and only preserves the component corresponding to the 1-d subspace
CΩ . Notice that the unit vector Ω is obtained from embedding the constant 1 ∈ C
via the map V , i.e., Ω = V 1. In matrix notation, if we identify C with its image CΩ
in K , then ϕ(A) is put into a matrix corner:
" #
ϕ(A) ∗
π(A) =
∗ ∗
Equivalently
ϕ(A) = P1 π(A) : P1 K → C;
ϕ(A) = V ∗ π(A)V, ∀A ∈ A.
Notice that this construction starts with a possibly infinite dimensional Hilbert space
H (instead of the 1-dimensional Hilbert space C), the map V embeds H into a
bigger Hilbert space K . If H is identified with its image in K , then π(A) is put
into a matrix corner, " #
π(A) ∗
∗ ∗
so that when acting on vectors,
" #" #
h i π(A) ∗ Vξ
ϕ(A)ξ = V ξ 0 .
∗ ∗ 0
ϕ(A) = PH π(A)
i.e., π(A) can be put into a matrix corner. K is chosen as minimal in the sense that
K = span{π(A)(V h) : A ∈ A, h ∈ H }.
H
V /K
ϕ(A) π(A)
H /K
V
5.3 Stinespring’s Theorem 151
Note: The containment H ⊂ K comes after the identification of H with its im-
age in K under the isometric embedding V . We write ϕ(A) = PH π(A), as
opposed to ϕ(A) = PH π(A)PH , since ϕ(A) only acts on the subspace H .
Theorem 5.1 (Stinespring [Sti55]). Let A be a ∗-algebra. The following are equiv-
alent:
W π1 = π2W (5.13)
Proof.
(Part 3, uniqueness) Let (Vi , Ki , πi ), i = 1, 2, be as in the statement satisfying
(5.11)-(5.12). Define
W π1 (A)V h = π2 (A)V h
Since such vectors are dense in the respective dilated space, we conclude that W π1 =
π2W , so (5.13) holds.
Note: kπ1 (B)π1 (A)V hk2 = hh,V ∗ π1 (A∗ B∗ BA)V hi. Fix A ∈ B(H ), the map B 7→
A∗ BA is an automorphism on B(H ).
(2 =⇒ 1)
Now suppose ϕ(A) = V ∗ π(A)V , and we verify it is completely positive.
Since positive elements in A ⊗ Mn are sums of the operator matrix
A∗1
∗ h
A2 i
∑ A∗i A j ⊗ ei j = .
.
A1 A2 · · · An
i, j .
A∗n
RHS(5.14) = ∑ vi , ϕ (A∗i A j ) v j H
i, j
5.3 Stinespring’s Theorem 153
= ∑ vi ,V ∗ π (A∗i A j )V v j H
i, j
= ∑ π (Ai )V vi , π (A j )V v j K
i, j
2
= ∑ π (Ai )V vi ≥ 0.
i K
(1 =⇒ 2)
Given a completely positive map ϕ, we construct K = Kϕ , V = Vϕ and
ϕ ⊗ IMn : A ⊗ Mn → B(H ⊗ Mn )
is positive, and
ϕ ⊗ IMn (1A ⊗ IMn ) = IH ⊗ IMn .
The condition on the identity element can be stated using matrix notation as
ϕ 0 ··· 0 1A 0 · · · 0 IH 0 · · · 0
0 ϕ ··· 0 0 1A · · · 0 0 IH · · · 0
. .. . . .. .. .. . . .. = .. .. . . .. .
.
. . . .
. . . . . . . .
0 0 ··· ϕ 0 0 · · · 1A 0 0 · · · IH
Let V : H → K0 , by
V ξ := 1A ⊗ ξ , ∀ξ ∈ H .
Then,
kV ξ k2ϕ = h1A ⊗ ξ , 1A ⊗ ξ iϕ
= hξ , ϕ(1∗A 1A )ξ iH
= hξ , ξ iH = kξ k2H
V
i.e., V is isometric, and so H −→ K0 is an isometric embedding.
Claim.
(i) V ∗V = IH ;
(ii) VV ∗ = projection from K0 on the subspace 1A ⊗ H .
Indeed, for any A ⊗ η ∈ K0 , we have
hA ⊗ η,V ξ iϕ = hA ⊗ η, 1A ⊗ ξ iϕ
= hη, ϕ(A∗ )ξ iH
= hϕ(A∗ )∗ η, ξ iH
It follows that
V ∗V ξ = V ∗ (1A ⊗ ξ ) = ϕ(1∗A )∗ ξ = ξ , ∀ξ ∈ H
VV ∗ (A ⊗ η) = V (ϕ(A∗ )∗ η) = 1A ⊗ ϕ(A∗ )∗ η.
5.3 Stinespring’s Theorem 155
This proves the claim. It is clear that the properties of V pass to the dilated space
K = Kϕ = clϕ (K0 /N).
To finish the proof of the theorem, define π = πϕ as follows: Set
!
π(A) ∑Bj ⊗ηj := ∑ AB j ⊗ η j , ∀A ∈ A
j j
and extend it to K .
For all ξ , η ∈ H , then,
N
∑ A∗i Ai = IH ; (5.16)
i=1
Now let (π, K ) be the pair obtained from Theorem 5.1 (Strinespring); then as a
block-matrix of operators, we have as follows
" #
∗ Ai ∗
π (si ) = (5.18)
0 ∗
5.4 Comments
In Stinespring’s theorem, the dilated space comes from a general principle (using
positive definite functions) when building Hilbert spaces out of the given data. We
illustrate this point with a few familiar examples.
Example 5.1. In linear algebra, there is a bijection between inner product structures
on Cn and positive-definite n × n matrices. Specifically, h·, ·i : Cn × Cn → C is an
inner product if and only if there exists a positive definite matrix A such that
hv, wiA = v∗ Aw
for all v, w ∈ Cn . We think of Cn as C-valued functions on {1, 2, . . . , n}, then h·, ·iA
is an inner product built on the function space.
2 * +
∑ ci δxi := ∑ ci δxi , ∑ c j δx j = ∑ ci c j F (xi , x j ) ≥ 0.
i F i j F i, j
K0 = span{δA : A ∈ M} = span{χA : A ∈ M}
Note the index set here is B (X), and ∑i ci δAi = ∑i ci χAi , i.e., these are precisely the
simple functions . Define
* +
∑ ci χAi , ∑ d j χB j := ∑ ci d j µ(Ai ∩ B j )
i j i, j
so that
2
∑ ci δAi = ∑ ci c j ϕ (A∗i A j ) ≥ 0.
i ϕ i, j
is indeed positive definite. But when extend linearly, one is in trouble. For
* +
∑ Ai ⊗ ξi , ∑ B j ⊗ η j
i j ϕ
=∑ ξi , ϕ(A∗i B j )η j H
i, j
and it is not clear why the matrix (ϕ(A∗i B j ))ni, j=1 should be a positive operator acting
in H ⊗ Cn . But we could very well put this extra requirement into an axiom, so the
CP condition (5.1).
ϕ(A) = V ∗ π(A)V
P2 = VV ∗VV ∗ = V (V ∗V )V ∗ = VV ∗ .
Summary
but not the other way around. In Nelson’s notes [Nel69], we use the notation ϕ ⊂ π
for one representation being the subrepresentation of another. To imitate the situa-
160 5 Completely Positive Maps
tion in linear algebra, we may want to split an operator T acting on K into operators
action on H and its complement in K . Let P : K → H be the orthogonal projec-
tion. In matrix language, " #
PT P PT P⊥
.
P⊥ T P P⊥ T P⊥
A better looking would be
" # " #
PT P 0 ϕ1 0
=
0 P⊥ T P⊥ 0 ϕ2
hence
π = ϕ1 ⊕ ϕ2 .
Stinespring’s theorem is more general, where the off-diagonal entries may not be
zero.
n times n times
z }| { z }| {
idH ⊕ · · · ⊕ idH ∈ Rep(A, H ⊕ · · · ⊕ H )
ηn Vn ξ
η1
* +
h i
∗
η2
= ∗ ∗
V1 V2 · · · Vn . ,ξ .
..
ηn
h i
This shows that V ∗ = V1∗ V2∗ · · · Vn∗ .
Corollary 5.2 (Krauss). Let dimH = n. Then all the CP maps are of the form
(The essential part here is that for any CP mapping ϕ, we get a system {Vi }.)
This was discovered in the physics literature by Kraus. The original proof was
very intricate, but it is a corollary of Stinespring’s theorem. When dimH = n, let
{e1 , . . . en } be an ONB. Fix a CP map ϕ, and get (V, K , π). Set
Vi : ei 7→ Vei ∈ K , i = 1, . . . n;
1. ⊕n1 H ' H ⊗ Cn
2. ∑⊕∞
1 H ' H ⊗l
2
3. Given L2 (X, M, µ), then L2 (X, H ) ' H ⊗ L2 (µ); where L2 (X, H ) con-
sists of all measurable functions f : X → H such that
ˆ
k f (x)k2H dµ(x) < ∞
X
and ˆ
h f , gi = h f (x) , g (x)iH dµ (x) .
X
For readers wishing to follow up sources, or to go in more depth with topics above,
we suggest: [Arv98, BR81, Sti59, Tak79].
Chapter 6
Brownian Motion
• Ω = sample space
• F = sigma algebra of events
• P = probability measure defined on F.
ˆ b
1 2 2
P ({a < X (ω) < b}) = √ e−(x−m) /2σ dx. (6.2)
a σ 2π
(The function under the integral in (6.2) is called the Gaussian distribution.)
165
166 6 Brownian Motion
P (A ∩ B) = P (A) P (B) .
Random variables X and Y are said to be independent iff (Def.) X −1 (I) and
Y −1 (J) are independent for all intervals I and J.
1. the random variables Xt1 , Xt2 , . . . , Xtn are jointly Gaussian with
ˆ
E (Xt ) = Xt (ω) dP (ω) = 0, ∀t ∈ R;
Ω
2. if t1 < t2 < · · · < tn then Xti+1 − Xti and Xti − Xti−1 are independent;
3. for all s,t ∈ R,
E |Xt − Xs |2 = |t − s| .
Exercise 6.1. Using Definition 6.3, Corollary 1.2, Remark 1.6 and Example 1.1,
show that there is a unique strongly continuous unitary one-parameter group {U (t)}t∈R
acting in L2 (C (R) , Cyl, P), determined by
Theorem 6.1 (see e.g., [Nel67]). Set Ω = C (R) = (all continuous real valued func-
tion on R), F = the sigma algebra generated by cylinder-sets, i.e., determined by
6.1 The Path Space 167
The measure P is determined by its value on cylinder sets, and and an integral over
Gaussians; it is called the Wiener-measure. Set
If 0 < t1 < t2 < · · · < tn , and the cylinder set is as in 6.5, then
where
1 −x2 /2t
gt (x) = √ e , ∀t > 0,
2πt
i.e., the N(0,t)-Gaussian. See Fig 6.1.
J2 gti -ti-1
gt1 Jn
gtn -tn-1
J1
Jn-1
Ji
t
t1 t2 ... ti ti+1 ... tn-1 tn
Fig. 6.1: Stochastic processes indexed by time: A cylinder set C is a special subset
of the space of all paths, i.e., functions of a time variable. A fixed cylinder set C is
specified by a finite set of sample point on the time-axis (horizontal), and a corre-
sponding set of “windows” (intervals on the vertical axis). When sample points and
intervals are given, we define the corresponding cylinder set C to be the set of all
paths that pass through the respective windows at the sampled times. In the figure
we illustrate sample points (say future relative to t = 0). Imagine the set C of all
outcomes with specification at the points t1 ,t2 , . . . etc.
168 6 Brownian Motion
I3 t
t1 t2 t3
I2
I1 '
I1
I2 '
I3 '
t
t1 t2 t3
(a) C optimistic (b) C0 pessimistic
I1 '' I2 ''
t
t1 t2 t3
I3 ''
Let Ω = ∏∞
k=1 {1, −1} be the infinite Cartesian product of {1, −1} with the prod-
uct topology. Ω is compact and Hausdorff by Tychnoff’s theorem.
For each k ∈ N, let Xk : Ω → {1, −1} be the kth coordinate projection, and assign
probability measures µk on Ω so that µk ◦ Xk−1 {1} = a and µk ◦ Xk−1 {−1} = 1 − a,
where a ∈ (0, 1). The collection of measures {µk } satisfies the consistency con-
dition, i.e., µk is the restriction of µk+1 onto the kth coordinate space. By Kolo-
mogorov’s extension theorem, there exists a unique probability measure P on Ω so
that the restriction of P to the kth coordinate is equal to µk .
It follows that {Xk } is a sequence of independent identically distributed (i.i.d.)
random variables in L2 (Ω , P) with E (Xk ) = 0 and Var[Xk2 ] = 1; and L2 (Ω , P) =
span{Xk }.
Remark 6.1. Let H be a separable Hilbert space with an orthonormal basis {uk }.
The map ϕ : uk 7→ Xk extends linearly to an isometric embedding of H into
L2 (Ω , P). Moreover, let F+ (H ) be the symmetric Fock space. F+ (H ) is the
closed span of the the algebraic tensors uk1 ⊗ · · · ⊗ ukn , thus ϕ extends to an iso-
morphism from F+ (H ) to L2 (Ω , P).
1 1
Exercise 6.4. Let Ω = ∏∞ 1 {−1, 1}, and let µ be the “fair-coin” measure 2 , 2 on
{±1} (e.g., “Head v.s. Tail”), let F be the cylinder sigma-algebra of subsets of Ω .
6.2 Decomposition of Brownian motion 169
∞ ˆ t
Xt (ω) := ∑ ψ j (s) ds Z j (ω) , t ∈ [0, 1] , ω ∈ Ω .
j=1 0
Show that the {Xt }t∈[0,1] is Brownian motion, where time “t” is restricted to [0, 1].
Hint: Let t1 ,t2 ∈ [0, 1], then
ˆ t1 ˆ t2 !
E (Xt1 Xt2 ) = E ∑ ψ j (s) ds Z j · ∑ ψk (s) ds Zk
j 0 k 0
ˆ t1 ˆ t2
= ∑∑ ψ j (s) ds ψk (s) dsE (Z j Zk )
j k 0 0 | {z }
=δ jk
= ∑ χ[0,t1 ] , ψ j L2
χ[0,t2 ] , ψ j L2
j
K(s,t) = s ∧ t
d2
− u= f
dx2
probability space (Ω , B, P), where the parameter t usually represents time. {Xt } is
a Brownian motion process if it is a mean zero Gaussian process such that
ˆ
E[Xs Xt ] = Xs Xt dP = s ∧ t.
Ω
Ω = ∏ R̄
t
Xt : Ω → R
is defined as
Xt (ω) = ω(t)
It it this property that makes the set of differentiable functions have measure zero.
In this sense, the trajectory of Brownian motion is nowhere differentiable.
An very important application of the spectral theorem of compact operators in to
decompose the Brownian motion process.
∞
sin(nπt)
Bt (ω) = ∑ Zn (ω)
n=1 nπ
where ∞
sin (nπs) sin (nπt)
s∧t = ∑
n=1 (nπ)2
and Zn ∼ N(0, 1).
Exercise 6.5. Look up the Central Limit Theorem (CLT), and prove the following
approximation formula for Brownian motion:
6.2 Decomposition of Brownian motion 171
1 n
Sn (·) = √ ∑ Wk (·) . (6.6)
n k=1
Let Xt denote Brownian motion. Then show that
1 bntc
Xt (·) = lim √ ∑ Wk (·) (6.7)
n→∞ n
k=1
We claim that
EP (Xs Xt ) = s ∧ t. (6.8)
1 bn sc bntc
= ∑ ∑ δ j,k
n j=1 k=1
bn sc
= → s, as n → ∞.
n
172 6 Brownian Motion
Xt Xt
t t
0.5 1 0.5 1
(a) n = 10 (b) n = 50
Xt Xt
t t
0.5 1 0.5 1
For readers wishing to follow up sources, or to go in more depth with topics above,
we suggest: [Hid80, Itô04, Itô06, Itô07, Nel67, Par09].
Chapter 7
Applications to Groups
— A.N. Kolmogorov
7.0.1 Motivation
g∗ := g−1 .
(g ⊗ cg )(h ⊗ ch ) = gh ⊗ cg ch
t(g ⊗ cg ) = g ⊗ tcg
173
174 7 Applications to Groups
(g ⊗ cg )∗ = g−1 ⊗ cg−1 .
Note: The ∗ operation so defined is the only way to make it a period-2, conjugate
linear, anti-automorphism. That is, the ∗-operation on AG is conjugate linear
and satisfies A∗∗ = A, and (AB)∗ = B∗ A∗ , for all A, B ∈ AG .
π̃(g ⊗ cg ) = πg ⊗ cg .
π =ρ G⊗1
is a representation of G ' G ⊗ 1.
Note: The notation g ⊗ cg is usually written as cg , which contains all the informa-
tion. Thus the pointwise multiplication takes the form
cg dh = lgh .
Equivalently,
lg = ∑ ch dh−1 g
h
More generally, every locally compact group has a left (and right) Haar measure.
Thus the above construction has a “continuous” version. It turns out that AG is
the Banach ∗-algebra L1 (G). Again, there is a bijection between Rep(G, H ) and
Rep(L1 (G), H ). In particular, the “discrete” version is recovered if the measure µ
is discrete, in which case AG = l 1 (G).
7 Applications to Groups 175
Everything we say about algebras is also true for groups. In physics, we are inter-
ested in representation of symmetry groups, which preserve inner product or energy.
We always want unitary representations. Irreducible representation amounts to ele-
mentary particles which can not be broken up further. In practice, quite a lot work
goes into finding irreducible representations of symmetry groups. The idea is to go
from groups to algebras and then to representations.
G→A→π
• πG ∈ Rep(G, H )
π(g1 g2 ) = π(g1 )π(g2 )
π(eG ) = IH
π(g)∗
= π(g−1 )
• πA ∈ Rep(A, H )
π(A1 A2 ) = π(A1 )π(A)2
π(1A ) = IH
π(A)∗
= π(A∗ )
!∗ !
∑ c(g)g = ∑ c(g−1 )g
g g
Existence of Haar measure: easy proof for compact groups, and extend to locally
compact cases. For non compact groups, the left / right Haar measures could be dif-
ferent. If they are always equal, the group is called unimodular. Many non compact
176 7 Applications to Groups
Let B(G) be the Borel σ -aglebra of G, E ∈ B(G). Left Haar: λL (gE) = λL (E);
right Haar: λR (Eg) = λR (Eg). Theorem: the two measures are equivalent
λL λR λL
dλL
4G = :G→G
dλR
is called the modular function, 4G is a homomorphism, 4G (gh) = 4G (g)4G (h).
L1 (G)
ˆ
(ϕ1 ? ϕ2 )(g) = ϕ1 (gh−1 )ϕ2 (h)dλR (h)
Remark 7.1. (1) for the case l 1 (G), the counting measure in unimodular, hence 4(g)
does not appear; (2) In ϕ1 (gh−1 ), h−1 appears since operation on points is dual to
operation on functions. A change of variable shows
ˆ ˆ ˆ
ϕ(gh−1 )dλR (g) = ϕ(g0 )dλR (g0 h) = ϕ(g0 )dλR (g0 )
(3) L1 (G) is a Banach algebra. Fubini’s theorem shows that f ? g ∈ L1 (G), for all
f , g ∈ L1 (G).
x 7→ ax + b
" #" #
1
− ba da db dadb
dλL = g−1 dg = a =
0 1 0 0 a2
check:
0 a b−b0
" # " # " #" # " #
ab a0 b0 −1
1
a0 − ba0 ab a0 a0
g= ,h= ,h g= =
01 0 1 0 1 01 0 1
178 7 Applications to Groups
ˆ ˆ
a b − b0 dadb
f (h−1 g)dλL (g) = f( 0, ) 2
a a0 a
ˆ
d(a0 s)d(a0t + b0 )
= f (s,t)
(a0 s)(a0 s)
ˆ
dsdt
= f (s,t) 2
s
a
s= , da = a0 ds
a0
b − b0
t= , db = a0 dt
a0
dadb a02 dsdt dsdt
2
= 0 2
= 2
a (sa ) s
Right Haar:
ˆ ˆ
f (gh−1 )(dg)g−1 = f (g0 )d(g0 h)(g0 h)−1
ˆ
= f (g0 )(dg0 )(hh−1 )g0−1
ˆ
= f (g0 )(dg0 )g0−1
" #" #
1
−1 da db a − ba dadb
dλR = (dg)g = =
0 0 0 1 a
check:
0 0
" # " # " #" # " #
ab a0 b0 ab 1
a0 − ab0 a
a0 − ab
a0 + b
g= ,h= , gh−1 = =
01 0 1 01 0 1 0 1
ˆ ˆ
−1 a ab0 dadb
f (gh )dλR (g) = f ( 0 , − 0 + b)
a a a
ˆ 0
a dsdt
= f (s,t) 0
as
ˆ
dsdt
= f (s,t)
s
ab0
t =− + b, db = dt
a0
dadb a0 dsdt dsdt
= =
a a0 s s
Most of the Lie groups considered here fall in the following class:
Let V be a finite-dimensional vector space over R (or C). We will consider the
real case here, but the modifications needed for the complex case are straightfor-
ward.
Let q : V × V → R be a non-degenerate form, i.e., bilinear s.t. v → q (v, ·) ∈ V ∗
is 1-1. Let GL (V ) be the general linear group for V , i.e., all invertible linear maps
V → V . (If a basis in V is chosen, this will be a matrix-group.)
Then G (q) is a Lie group, and its Lie algebra consists of all linear mapping X : V →
V such that
q (Xu, v) + q (u, Xv) = 0, ∀u, v ∈ V. (7.2)
i.e., the matrix-exponential. Note that gX (t) satisfies (7.1) for all t ∈ R iff X satisfies
(7.2). To see this, differentiate: i.e., compute
d
q (exp (tX) , exp (tX) v)
dt
• ax + b
• Heisenberg
• SL2 (R)
• Lorentz
• Poincaré
Among these, the ax + b, Heisenberg and Poincaré groups are semi-direct prod-
uct groups. Their representations are induced from normal subgroups. It is ex-
tremely easy to find representations of abelian subgroups. Unitary representation
of abelian subgroups are one-dimensional, but the induced representation on an en-
larged Hilbert space is infinite dimensional. See the appendix for a quick review of
semi-direct product.
" #
ab
Example 7.1. The ax + b group (a > 0). G = {(a, b)}, where (a, b) = . The
01
multiplication rule is given by
check:
Example 7.3. Form L2 (µL ) where µL is the left Haar measure. Then π : g →
π(g) f (x) = f (g−1 x) is a unitary representation. Specifically, if g = (a, b) then
x y−b
f (g−1 x) = f ( , ).
a a
d x y−b ∂ ∂
X̃ f = a=1,b=0
f( , ) = (−x − y ) f (x, y)
da a a ∂x ∂y
d x y−b ∂
Ỹ f = a=1,b=0
f( , ) = − f (x, y)
db a a ∂y
∂ ∂
X̃ = −x −y
∂x ∂y
∂
Ỹ = −
∂y
[X̃, Ỹ ] = X̃ Ỹ − Ỹ X̃
∂ ∂ ∂ ∂ ∂ ∂
= (−x − y )(− ) − (− )(−x − y )
∂x ∂y ∂y ∂y ∂x ∂y
∂2 ∂2 ∂2 ∂ ∂2
=x + y 2 − (x + +y 2)
∂ x∂ y ∂y ∂ x∂ y ∂ y ∂y
∂
=−
∂y
= Ỹ .
d
X̃ f = t=0
f (e−tX x)
dt
d
= t=0
f ((e−t , 1)(x, y))
dt
d
= t=0
f (e−t x, e−t y + 1)
dt
∂ ∂
= (−x − y ) f (x, y)
∂x ∂y
d
Y˜f = t=0
f (e−tY x)
dt
d
= t=0
f ((1, −t)(x, y))
dt
d
= t=0
f (x, y − t)
dt
∂
=− f (x, y)
∂y
Example 7.4. We may parametrize the Lie algebra of the ax+b group using (x, y)
variables. Build the Hilbert space L2 (µL ). The unitary representation π(g) f (σ ) =
f (g−1 σ ) induces the follows representations of the Lie algebra
d
dπ(s) f (σ ) = s=0
f (e−sX σ ) = X̃ f (σ )
dx
7.2 Induced representation 183
d
dπ(t) f (σ ) = t=0
f (e−tY σ ) = Ỹ f (σ ).
dy
Hence in the parameter space (s,t) ∈ R2 we have two usual derivative operators
∂ /∂ s and ∂ /∂t, where on the manifold we have
∂ ∂ ∂
= −x − y
∂s ∂x ∂y
∂ ∂
=−
∂t ∂y
has trace trace = x2 + y2 + 1 ≥ 1, and det = x2 ≥ 0. If instead we have “y2 ” then the
determinant is the constant zero.
Example 7.5. Notice the term “y2 + 1” is essential for 4 being elliptic. Also notice
that all the coefficients are analytic functions in the (x, y) variables.
Note: Notice the Γ is unimodular, hence it is just a copy of R. Its invariant measure
is the Lebesgue measure on R.
001
001
ϕ : (R, +) → Aut(Γ )
" # " #" #
c 1a c
ϕ(a) =
b 01 b
" #
c + ab
=
b
check:
This also goes under the name of “Mackey machine”. Its modern formulation is in
the context of completely positive map.
Let G be a locally compact group, and Γ ⊂ G be a closed subgroup. Recall the
modular functions come in when the translation was put on the wrong side, i.e.
ˆ ˆ
−1
f (gx)dx = 4(g ) f (x)dx
G G
or equivalently, ˆ ˆ
4(g) f (gx)dx = f (x)dx.
G G
Note: Since ϕ has compact support, the integral is well-defined. τ is called condi-
tional expectation. It is simply the summation of ϕ over the orbit Γ x. This
is because if ξ runs over Γ , ξ x runs over Γ x. τϕ may also be interpreted as
taking average, only it does not divide out the total mass, but that only differs
by a constant.
τϕ(ξ x) = τϕ(x), ∀ξ ∈ Γ .
(τϕ)(z0 + x) = ∑ ϕ(z0 + z + x)
z∈Z
= ∑ ϕ(z + x)
z∈Z
Since ϕ has compact support, ϕ(z + x) vanishes for all but a finite number of z.
Hence τϕ contains a finite summation, and so it is well-defined.
(Rg f )(ξ x) = f (ξ xg) = ρ(ξ )1/2 Lξ f (xg) = ρ(ξ )1/2 Lξ (Rg f )(x)
so that Rg f ∈ F∗ .
Note: The factor ρ(ξ )1/2 comes in, since later we will defined an inner product on
F∗ so that k f (ξ ·)knew = k f (·)knew , ∀ξ ∈ Γ . Eventually, we will define the
induced representation by (Ugind f )(·) := (Rg f )(·), not on F∗ , but pass to a
quotient space.
Note: Let’s ignore the factor ρ(ξ )1/2 for a moment. Lξ is unitary implies
that for all f ∈ F∗ ,
dadb
dλR =
a
dadb
dλL =
a2
dλL 1
4= =
dλR a
188 7 Applications to Groups
is a positive linear functional. By Riesz’s theorem, there exists a unique Radon mea-
sure µ f , f on M, such that
ˆ ˆ
k f (g)kV2 ϕ(g)dg = (τϕ)dµ f , f .
G M
Note: Recall that given a measure space (X, M, µ), let f : X → Y . Define a linear
functional Λ : Cc (Y ) → C by
ˆ
Λ ϕ := ϕ( f (x))dµ(x)
It follows that µ f = µ ◦ f −1 .
Ugind is unitary,
Ugind f = k f kH .
H
´
Note: µ f ,g (M) = M τϕdξ , where τϕ ≡ 1. What is ϕ then? It turns out that ϕ could
be constant 1 over a single Γ -period, or ϕ could spread out to a finite number
of Γ -periods. Therefore, in this case
ˆ
2
kfk = k f (g)kV2 ϕ(g)dg
ˆ G
= k f (g)kV2 ϕ(g)dg
1−period
ˆ
= k f (g)kV2 dg
1−period
ˆ
= k f (g)kV2 dg
M
Proof. Check:
Ug P(ψ)Ug−1 = P(ψ(·g)),
001
001
i.e. Ad : G → GL(n), as
7.3 Example - Heisenberg group 191
Ad(g)(n) : = gng−1
(x, y) 7→ (ax + y)
Ug f (x) = eih(c+bx) f (x + a)
Lξ (b,c) = eihc .
Since
f (x, y, z) satisfies
i.e. we may translate the y, z variables by arbitrary amount, and the only
price to pay is multiplicative factor eihc . Therefore f is really a function
defined on the quotient
M = Γ \G ' R.
M = {(x, 0, 0)} is identified with R, and the invariant measure on the ho-
mogeneous space M is simply the Lebesgue measure. It is almost clear
192 7 Applications to Groups
where
ˆ
(τϕ)(π(g)) = ϕ(ξ g)dξ
ˆΓ
= ϕ((0, b, c)(x, y, z))dbdc
2
ˆR
= ϕ(x, b + y, c + z)dbdc
2
ˆR
= ϕ(x, b, c)dbdc
2
ˆR
= ϕ(x, y, z)dydz
R2
= (τϕ)(x).
Λ = µf,f
7.3 Example - Heisenberg group 193
i.e. ˆ ˆ
2
| f (x, y, z)| ϕ(x, y, z)dxdydz = (τϕ)(x)dµ f , f (x).
R3 R
4. Define
ˆ ˆ ˆ
k f k2ind := µ f , f (M) = 2
| f | dξ = 2
| f (x)| dx = | f (x, 0, 0)|2 dx
M R R
Ugind f (g0 ) := f (g0 g)
W : H ind → L2 (R)
(W f )(x) = f (x, 0, 0)
If put other numbers into f , as f (x, y, z), the result is the same, since
f ∈ H ind is really defined on the quotient M = Γ \G ' R.
W is unitary:
ˆ ˆ ˆ
kW f k2L2 = 2
|W f | dx = 2
| f (x, 0, 0)| dx = | f |2 dξ = k f k2ind
R R Γ \G
Ug (W f ) = eih(c+bx) f (x + a, 0, 0)
WUgind f = W ( f ((x, y, z)(a, b, c)))
= W ( f (x + a, y + b, z + c + xb))
= W eih(c+bx) f (x + a, y, z)
= eih(c+bx) f (x + a, 0, 0)
194 7 Applications to Groups
6. Since {U, L}0 ⊂ {L}0 , then the system {U, L} is reducible implies L is re-
ducible. Equivalent, {L} is irreducible implies {U, L} is irreducible. Con-
sequently, Ug is irreducible. Since Ug is the induced representation, if it is
reducible, L would also be reducible, but L is 1-dimensional.
Note: The Heisenberg group is a non abelian unimodular Lie group, so the Haar
measure on G is just the product measure dxdydz on R3 . Conditional expec-
tation becomes integrating out the variables correspond to the subgroup. For
example, given f (x, y, z) conditioning with respect to the subgroup (0, b, c)
amounts to integrating out the y, z variable and get a function f˜(x), where
¨
f˜(x) = f (x, y, z)dydz, i.e.,
Γ \G ' R.
7.3.1 ax + b group
" #
ab
a ∈ R+ , b ∈ R, g = (a, b) = .
01
Ug f (x) = eiax f (x + b)
t
Ug f (x) = eie x f (x + b)
x f (x+b)
Ug(a,b) f (x) = eiae
d
[ , iex ] = iex
dx
[A, B] = B
or
x
U(et ,b) f = eite f (x + b)
" #
0b
01
7.4 Co-adjoint Orbits 195
It turns out that only a small family of representations are induced. The question is
how to detect whether a representation is induced. The whole theory is also under
the name of “Mackey machine” [Mac52, Mac88]. The notion of “machine” refers
to something that one can actually compute in practice. Two main examples are the
Heisenberg group and the ax + b group.
What is the mysteries parameter h that comes into the Schrödinger representa-
tion? It is a physical constant, but how to explain it in mathematical theory?
000
All the Lie groups the we will ever encounter come from a quadratic form. Given a
quadratic form
ϕ : V ×V → C
there is an associated group that fixes ϕ, i.e. we consider elements g such that
and define G(ϕ) as the collection of these elements. G(ϕ) is clearly a group. Apply
the exponential map and the product rule,
d
t=0
ϕ(etX x, etX y) = 0 ⇐⇒ ϕ(Xx, y) + ϕ(x, Xy) = 0
dt
196 7 Applications to Groups
hence
X + X tr = 0
det(etX ) = et·trace(X)
thus det = 1 if and only if trace = 0. It is often stated in differential geometry that
the derivative of the determinant is equal to the trace.
Example 7.10. Rn , ϕ(x, y) = ∑ xi yi . The associated group is the orthogonal group
On .
There is a famous cute little trick to make On−1 into a subgroup of On . On−1 is not
normal in On . We may split the quadratic form into
n−1
∑ xi2 + 1
i=1
In = {g : gu = u} ' On−1
Notice that for all v ∈ Sn−1 , there exists g ∈ On such that gu = v. Hence
g 7→ gu
On /On−1 ' Sn
Other examples of homogeneous spaces show up in number theory all the time. For
example, the Poincaré group G/discrete subgroup.
G, N ⊂ G normal subgroup. The map g·g−1 : G → G is an automorphism sending
identity to identity, hence if we differentiate it, we get a transformation in GL(g). i.e.
we get a family of maps Adg ∈ GL(g) indexed by elements in G. g 7→ Adg ∈ GL(g)
is a representation of G, hence if it is differentiated, we get a representation of g,
adg : g 7→ End(g) acting on the vector space g.
gng−1 ∈ N. ∀g, g · g−1 is a transformation from N to N, define Adg (n) = gng−1 .
Differentiate to get ad : n → n. n is a vector space, has a dual. Linear transformation
on vector space passes to the dual space.
In order to get the transformation rules work out, have to pass to the adjoint or the
dual space.
Adg∗ : n∗ → n∗
001
001
001
Adg : n → n given by
We use [ξ , η]T for the dual n∗ ; and use [x, y]T for n. Then
" # " #
ξ ξ
Adg∗ : 7→
η aξ + η
What about the orbit? In the example of On /On−1 , the orbit is Sn−1 .
For ξ ∈ R\{0}, the orbit of Adg∗ is
" # " #
ξ ξ
7→
0 R
i.e. vertical lines with x-coordinate ξ . ξ = 0 amounts to fixed point, i.e. the orbit is
a fixed point.
The simplest orbit is when the orbit is a fixed point. i.e.
7.5 Gårding Space 199
" # " #
ξ ξ
Adg∗ : 7→ ∈ V∗
η η
i.e. get vertical lines indexed by the x-coordinate ξ . In this example, a cross section
is a subset of R2 that intersects each orbit at precisely one point. Every cross section
in this example is a Borel set in R2 .
We don’t always get measurable cross sections. An example is the construction
of non-measurable set as was given in Rudin’s book. Cross section is a Borel set
that intersects each coset at precisely one point.
Why does it give all the equivalent classes of irreducible representations? Since
we have a unitary representation Ln ∈ Rep(N,V ), Ln : V → V and by construction
of the induced representation Ug ∈ Rep(G, H ), N ⊂ G normal such that
Ug LnUg−1 = Lgng−1
i.e.
Lg ' Lgng−1
Ln → LA → LA∗ .
U (exp (tX)) v − v
dU (X) v = lim
t→0 t
then
HGårding ⊂
\
dom (dU (X))
X∈g
and
dU (X) U (ϕ) v = U Xϕ
e v,
e (g) = d
Xϕ t=0
ϕ (exp (−tX) g) , ∀g ∈ G.
dt
Proof. (Hint)
ˆ ˆ
ϕ (g) U (exp (tX)) U (g) dg = ϕ (exp (−tX) g) U (g) dg.
G G
Almost all Lie algebras we will encounter come from specifying a quadratic form
ϕ : G × G → C. ϕ is then uniquely determined by a Hermitian matrix A so that
7.5 Gårding Space 201
ϕ(x, y) = xtr · Ay
d
t=0
ϕ(etX x, etX y) = 0
dt
X tr A + AX = 0
hence
g = {X : X tr A + AX = 0}.
for some selfadjoint operator HX (possibly unbounded). The RHS in (7.3) is given
by the Spectral Theorem. We often write
dU(X) := iHX
to indicate that dU(X) is the directional derivative along the direction X. Notice that
HX∗ = HX but
(iHX )∗ = −(iHX )
d d
t=0
U(etX1 ) f (x) = f (x)
dt dx
d
dU(X1 ) =
dx
202 7 Applications to Groups
d
U(etX2 ) f (x) = ihx f (x)
t=0
dt
dU(X2 ) = ihx
d
t=0
U(etX3 ) f (x) = ih f (x)
dt
dU(X2 ) = ihI
d
[dU(X1 ), dU(X2 )] = [ , ihx]
dx
d
= ih[ , x]
dx
= ih
1 d
−idU(X1 ) =
i dx
−idU(X2 ) = hx
−idU(X1 ) = hI
1 d h
[ , hx] = .
i dx i
Below we answer the following question:
What is the space of functions that Ug acts on? L. Gårding /gor-ding/ (Swedish
mathematician) looked for one space that always works. It’s now called the Gårding
space.
Start with Cc (G), every ϕ ∈ Cc (G) can be approximated by the so called Gårding
functions, using the convolution argument. Define convolution as
ˆ
ϕ ? ψ(g) = ϕ(gh)ψ(h)dR h
ˆG
ϕ ? ψ(g) = ϕ(h)ψ(g−1 h)dL h
G
ϕ ? ζ j → ϕ, j → 0.
Ζj
e G
Fig. 7.1: Approximation of identity.
Since ϕ vanishes outside a compact set, and since U(h)v is continuous and bounded
in k·k, it follows that U(ϕ) is well-defined.
Every representation U of a Lie group G induces a representation (also denote
U) of the group algebra:
Lemma 7.5. U(ϕ1 ?ϕ2 ) = U(ϕ1 )U(ϕ2 ) (U is a representation of the group algebra)
is dense in H .
dU(X)U(ϕ)v = U(X̃ϕ)v
1
(U(etX ) − I)U(ϕ)v = U(X̃ϕ)v.
lim
t→0 t
= ϕ(h)U(gh)dh
ˆG
= 4(g)ϕ(g−1 h)U(h)dh
G
set g = etX .
Note: If assuming unimodular, 4 does not show up. Otherwise, 4 is some cor-
rection term which is also differentiable. X̃ acts on ϕ as X̃ϕ. X̃ is called the
derivative of the translation operator etX .
Exercise 7.1. Show that Schwartz space S is the Gårding space for the Schrödinger
representation.
• G = T, Ĝ = Z
χn (z) = zn
χn (zw) = zn wn = χn (z)χn (w)
• G = R, Ĝ = R
χt (x) = eitx
2πkl
χl (k) = ei n
Ĝ is also a group, with group operation defined by (χ1 χ2 )(g) := χ1 (g)χ2 (g). Ĝ is
called the group characters.
Theorem 7.4 (Pontryagin). If G is a locally compact abelian group, then G ' Ĝˆ
ˆ where “'” means “natural iso-
(isomorphism between G and the double dual Ĝ,)
morphism.”
Note: This result first appears in 1930s in the annals of math, when John von Neu-
mann was the editor of the journal at the time. The original paper was hand
written. von Neumann rewrote it, since then the theorem became very popu-
lar, see [Rud90].
There are many groups that are not abelian. We want to study the duality question
in general. Examples:
• compact group
• finite group (abelian, or not)
• H3 locally compact, nonabelian, unimodular
• ax+b locally compact, nonabelian, non-unimodular
206 7 Applications to Groups
Then Rg is a unitary operator acting on L2 (µR ), where µR is the right invariant Haar
measure.
Theorem 7.5 (Krein, Weil, Segal). Let G be locally compact unimodular (abelian
or not). Then the right regular representation decomposes into a direct integral of
irreducible representations
ˆ ⊕
Rg = ”irrep” dµ
Ĝ
(Uy f )(x) = f (x + y)
= ∑ fˆ(n)χn (x + y)
n
= ∑ fˆ(n)ei2πn(x+y)
n
(Uy f )(x) = f (x + y)
ˆ
= fˆ(t)χt (x + y)dt
ˆ R
= fˆ(t)eit(x+y) dt
R
ˆ
(Uy f )(0) = f (y) = fˆ(t)eity dt
R
7.6 Decomposition of representation 207
As can be seen that Fourier series and Fourier integrals are special cases of the de-
composition of the right regular representation Rg of a unimodular locally compact
´⊕
group. =⇒ k f k = k fˆk. This is a result that was done 30 years earlier before
the non abelian case. Classical function theory studies other types of convergence,
pointwise, uniform, etc.
The duality question may also be asked for discrete subgroups. This leads to re-
markable applications in automorphic functions, automorphic forms, p-adic num-
bers, compact Riemann surface, hyperbolic geometry, etc.
Example 7.18. Cyclic group of order n. G = Z/nZ ' {0, 1, · · · , n − 1}. Ĝ = G. This
is another example where the dual group is identical to the group itself. Let ζ =
ei2π/n be the primitive nth -root of unity. k ∈ Zn , l = {0, 1, . . . , n − 1}
208 7 Applications to Groups
2πkl
χl (k) = ei n
In this case, Segal’s theorem gives finite Fourier transform. U : l 2 (Z) → l 2 (Ẑ) where
1
U f (l) = √ ∑ ζ kl f (k)
N k
Exercise 7.2. (i) Find the normal subgroups in the Heisenberg group. (ii) Find the
normal subgroups in the ax + b group.
May use this to induce a representation of G. This is called principle series. Need to
do something else to get all irreducible representations.
A theorem by Iwasawa states that simple matrix group (Lie group) can be de-
composed into
G = KAN
where K is compact, A is abelian and N is nilpotent. For example, in the SL2 case,
! ! !
cost − sint es 0 1u
SL2 (R) = .
sint cost 0 e−s 01
The simple groups do not have normal subgroups. The representations are much
more difficult.
χt (ν) = eitν
W Ft (h) = Ft (h).
So what does the induced representation look like in L2 (H) then? Recall by
definition that
Ut (g) := W indχGt (g) W ∗
indχGt
Ht / Ht
W W
L2 (H)
Ut
/ L2 (H)
Let f ∈ L2 (H).
Ut (g) f (h) = W indχGt (g) W ∗ f (h)
= indχGt (g)W ∗ f (h)
= (W ∗ f ) (hg) .
The Mackey machine does not cover many important symmetry groups in
physics. Actually most of these are simple groups. However it can still be
applied. For example, in special relativity theory, we have the Poincaré group
L n R4 where R4 is the normal subgroup. The baby version of this is when
L = SL2 (R). V. Bargman formulated this baby version. Wigner pioneered
the Mackey machine, long before Mackey was around.
Once we get unitary representations, differentiate it and get selfadjoint alge-
bra of operators (possibly unbounded). These are the observables in quantum
mechanics.
Define indχRt (y) f (x) = f (x + y). Claim that Ht ' L2 [0, 1]. The unitary transforma-
tion is given by W : Ht → L2 [0, 1]
(W Ft )(x) = Ft (x).
Let’s see what indχRt (y) looks like on L2 [0, 1]. For any f ∈ L2 [0, 1],
W indχGt (y) W ∗ f (x) = indχGt (y)W ∗ f (x)
= (W ∗ f ) (x + y)
Note: Are there any functions in Ht ? Yes, for example, f (x) = eitx . If f ∈ Ht , | f |
is 1-periodic. Therefore f is really a function defined on Z\R ' [0, 1]. Such
a function has the form
In Nelson’s notes [Nel69], a normal representation has the form (counting multi-
plicity)
⊕
ρ = ∑ nπ Hn
, Hn ⊥ Hm
where
nπ = π ⊕ · · · ⊕ π (n times)
⊕
∑ K = lZ2 n ⊗ K.
In matrix form, this is a diagonal matrix with π repeated on the diagonal n times. n
could be 1, 2, . . . , ∞. We apply this to group representations.
Locally compact group can be divided into the following types.
• abelian
• non-abelian: unimodular, non-unimodular
• non-abelian: Mackey machine, semidirect product e.g. H3 , ax + b; simple
group SL2 (R). Even it’s called simple, ironically its representation is much
more difficult than the semidirect product case.
We want to apply these to group representations.
Spectral theorem says that given a normal operator A, we may define f (A) for
quite a large class of functions, actually all measurable functions. One way to define
f (A) is to use the multiplication version of the spectral theorem, and let
f (A) = F f (Â)F −1 .
7.8 Connections to Nelson’s Spectral Theory 213
The other way is to use the projection-valued measure version of the spectral theo-
rem, write
ˆ
A= λ P(dλ )
ˆ
f (A) = f (λ )P(dλ ).
ρ : f 7→ ρ( f ) = f (A)
ρ( f g) = ρ( f )ρ(g)
Example 7.21. G = (R, +), group algebra L1 (R). Define Fourier transform
ˆ
fˆ(t) = f (x)e−itx dx.
Example 7.22. Fix t, H = C, ρ(·) = eit(·) ∈ Rep(G, H ). From the group represen-
tation ρ, we get a group algebra representation ρ̃ ∈ Rep(L1 (R), H ) defined by
ˆ ˆ
ρ̃( f ) = f (x)ρ(x)dx = f (x)eitx dx
It follows that
fˆ(ρ) := ρ̃( f )
? g = fˆĝ
? g = fd
fd
ρ(y) f (x) := f (x + y)
Define
fˆ(ρ) := ρ̂( f )
If we have used the left regular representation, instead of the right, then
ˆ
fˆ(ρ)g = ρ̃( f )g = f (y)ρ(y)g(·)dy
ˆ
= f (y)(Ly g)(·)dy
ˆ
= f (y)g(· − y)dy.
Back to the general case. Given a locally compact group G, form the group algebra
L1 (G), and define the left and right convolutions as
ˆ ˆ
(ϕ ? ψ)(x) = ϕ(g)ψ(g−1 x)dL g = ϕ(g)(Lg ψ)dL g
ˆ ˆ
(ϕ ? ψ)(x) = ϕ(xg)ψ(g)dR g = (Rg ϕ)ψ(g)dR g
and write
ψ̂(ρ) := ρ̃(ψ).
and
ˆ ˆ
ρ̃(ψ)ϕ = ψ(g)ρ(g)ϕdg = ψ(g)(Rg ϕ)dg
ˆG G
= ψ(g)ϕ(xg)dg
G
= (ϕ ? ψ)(x)
ρh : G → L2 (R)
where H is the normal subgroup {b, c}. It is not so nice to work with indHG (χh )
directly, so instead, we work with the equivalent representations, i.e. Schrödinger
representation. See Folland’s book on abstract harmonic analysis.
ˆ
ψ̂(h) = ψ(g)ρh (g)dg
G
Here the ψ̂ on the right hand side in the Fourier transform of ψ in the usual sense.
Therefore the operator ψ̂(h) is the one so that
4n ψ ∈ L1 (G), ψ ∈ C∞ (G).
d
t=0
(RetX ψ) = X̃ψ
dt
where ˆ
v= ϕ(g)ρ(g)wdg = ρ(ϕ)w. generalized convolution
If ρ = R, the n ˆ
v= ϕ(g)R(g)
X̃(ϕ ? w) = (Xϕ) ? w.
Example 7.25. H3
∂
a→
∂a
∂
b 7→
∂b
∂
c 7→
∂c
get standard Laplace operator. {ρh (ϕ)w} ⊂ L2 (R) . ” = ” due to Dixmier. {ρh (ϕ)w}
is the Schwartz space.
2 2
d d
+ (ihx)2 + (ih)2 = − (hx)2 − h2
dx dx
Notice that 2
d
− + (hx)2 + h2
dx
is the Harmonic oscillator. Spectrum = hZ+ .
For readers wishing to follow up sources, or to go in more depth with topics above,
we suggest: [JÓ00, Mac52, Mac85, Mac92].
Chapter 8
The Kadison-Singer Problem
The Kadison-Singer problem (KS) lies at the root of how questions from quantum
physics take shape in the language of functional analysis, and algebras of operators.
A brief sketch is included below, summarizing some recent advances (in fact the
KS-problem was recently solved.) It is of special interest as it is known that the
solution to KS at the same time answers a host of other questions; these with appli-
cations to engineering, especially to signal processing. The notion from functional
analysis here is “frame.” A frame of vectors in Hilbert space generalizes the notion
of orthonormal basis in Hilbert space.
The Kadison-Singer problem (KS) comes from functional analysis, but it was
resolved (only recently) with tools from areas of mathematics quite disparate from
functional analysis. More importantly, the solution to KS turned out to have impor-
tant implications for a host of applied fields from engineering.1
This reversal of the usual roles seem intriguing for a number of reasons: While
the applications considered so far, involve problems which in one way or the
other, derive from outside functional analysis itself, e.g., from physics, from signal-
processing, or from anyone of a number of areas of analysis, PDE, probability, statis-
tics, dynamics, ergodic theory, prediction theory etc.; the Kadison-Singer problem is
different. It comes directly from the foundational framework of functional analysis;
more specifically from the axiomatic formulation of C∗ -algebras. Then of course,
219
220 8 The Kadison-Singer Problem
Lemma 8.1. Pure states on B(H ) are unit vectors (in fact, the equivalent class of
unit vectors; equivalently, pure states sit inside the projective vector space. In Cn+1 ,
this is CPn .) Specifically, let u ∈ H , kuk = 1, then
2P.A.M. Dirac gave a lecture at Columbia University in the late 1950’s, in which he claimed
without proof that pure states on the algebra of diagonal operators (' l ∞ ) extends uniquely on
B (l 2 ). Two students Kadison and Singer sitting in the audience were skeptical about whether
Dirac knew what it meant to be an extension. They later formulated the conjecture in a joint paper
8 The Kadison-Singer Problem 221
theorem doesn’t guarantee the extension remains a pure state. Let E(s) be the set
of all states on B(H ) which extend s. E(s) is non-empty, compact and convex in
the weak-∗ topology. By Krein-Milman’s theorem, E(s) = closure(Extreme Points).
Any extreme point will then be a pure state extension of s; but which one to choose?
It’s the uniqueness part that is the famous KS problem.
Physics Mathematics
H ∗ = H, v ∈ H , kvk = 1 (state).
Spectral theorem: H ∼ PH (·)
projection-valued measure.
Measurement:
Prob (H ∈ (a, b)) =
kPH (a, b) vk2
Fig. 8.1: Observable, state, measurement. Left column: An idealized physics exper-
iment. Right: the mathematical counterpart, a selfadjoint operator H, its associated
projection-valued measure PH , and a norm-one vector v in Hilbert space.
β N 3 G 7−→ F + G ∈ β N
lim xn = ϕF (x) ;
F
λ1 ≥ λ2 ≥ · · · , λk = λk (A)
and set
n
1
trDix,ω (A) = lim ∑ λk (A) . (8.2)
ω log (n + 1) k=1
Definition 8.1. We say that A has finite Dixmier trace if the limit in (8.2) is finite.
The proof of the KS-problem involves systems of vectors in Hilbert space called
frames. For details we refer to [Cas13].
Below a sketch. Let H be a separable Hilbert space, and let {uk }k∈N be an ONB,
then we have the following unique representation
8.1 Frames in Hilbert Space 223
the Parseval-formula.
Definition 8.2. A system {vk }k∈N in H is called a frame if there are constants A, B
s.t. 0 < A ≤ B < ∞, and
Note that (8.5) generalizes (8.4). Below we show that, for frames, there is also a
natural extension of (8.3).
Proposition 8.1. Let {vk }k∈N be a frame in H ; then there is a dual system v∗k
k∈N
⊂
H such that the following representation holds:
Proof. Define the following operator T : H → l 2 (N) by Tw = (hvk , wiH )k∈N , and
show that the adjoint T ∗ : l 2 (N) → H satisfies T ∗ ((xk )) = ∑k∈N xk vk . Hence
Remark 8.2. We saw in Chapter 3 (sect. 3.3) that, if {vk }k∈N is an ONB in some
fixed Hilbert space H , then
Exercise 8.3. Write down the modified list of properties for P (·) in (8.8) which
generalize the axioms of Definition 3.4 for PVMs.
For readers wishing to follow up sources, or to go in more depth with topics above,
we suggest: [Arv76, BR81, Cas13, Cas14, AW13, MSS13].
The most current paper concerning the solution to KS appears to be [MSS13]
by Marcus, Spielman, and Strivastava. Paper [Cas14] by P. Casazza explains the
problem and its implications.
Part IV
Extension of Operators
Chapter 9
Selfadjoint Extensions
— Niels Bohr
227
228 9 Selfadjoint Extensions
Since the spectral theorem serves as the central tool in quantum measurements,
we must be precise about the distinction between linear operators with dense do-
main which are only Hermitian (formally selfadjoint) as opposed to selfadjoint.
This distinction is accounted for by von Neumann’s theory of deficiency indices
[AG93, DS88c, HdSS12]1 .
In order to apply spectral theorem, one must work with self adjoint operators in-
cluding the unbounded ones. Some examples first.
In quantum mechanics, to understand energy levels of atoms and radiation, the
energy level comes from discrete packages. The interactions are given by Column
Law where
c jk
H = −4r +
kr j − rk k
and Laplacian has dimension 3 × #(electrons).
In Schrödinger’s wave mechanics, one needs to solve for ψ(r,t) from the equa-
tion
1 ∂
Hψ = ψ.
i ∂t
If we apply spectral theorem, then ψ(t) = eitH ψ(r,t = 0). This shows that motion in
quantum mechanics is governed by unitary operators. The two parts in Schrödinger
equation are separately selfadjoint, but justification of the sum being selfadjoint
wasn’t made rigorous until 1957 when Kato wrote the book on “perturbation theory”
[Kat95]. It is a summary of the sum of selfadjoint operators.
In Heisenberg’s matrix mechanics, he suggested that one should look at two
states and the transition probability between them, such that
1 Starting with [vN32a, vN32d, vN32c], J. von Neumann and M. Stone did pioneering work in the
1930s on spectral theory for unbounded operators in Hilbert space; much of it in private correspon-
dence. The first named author has from conversations with M. Stone, that the notions “deficiency-
index,” and “deficiency space” are due to them; suggested by MS to vN as means of translating
more classical notions of “boundary values” into rigorous tools in abstract Hilbert space: closed
subspaces, projections, and dimension count.
9.1 Extensions of Hermitian Operators 229
If ψ(t) = eitH ψ, then it works. In Heisenberg’s picture, one looks at evolution of the
observables e−itH AeitH . In Schrödinger’s picture, one looks at evolution of states.
The two point of views are equivalent.
Everything so far is based on application of the spectral theorem, which requires
the operators being selfadjoint in the first place.
von Neumann’s index theory gives a complete classification of extensions of sin-
gle Hermitian unbounded operators with dense domain in a given Hilbert space. The
theory may be adapted to Hermitian representations of ∗-algebras [Nel59a].
Let A be a densely defined Hermitian operator on a Hilbert space H , i.e. A ⊂ A∗ .
If B is any Hermitian extension of A, then
A ⊂ B ⊂ B∗ ⊂ A∗ . (9.1)
e = A∗
A e) , where (9.4)
D (A
D A
e = D (A) ⊕ S
(9.5)
e ⊃ A.
defines a (closed) Hermitian operator A
230 9 Selfadjoint Extensions
Lemma 9.1. Let S be a closed subspace in D (A∗ ), where D (A∗ ) is a Hilbert space
under the A∗ -norm. The following are equivalent.
1. hA∗ y, xi = hy, A∗ xi, for all x, y ∈ S.
2. hx, A∗ xi ∈ R, for all x ∈ S.
Proof. If (1) holds, setting x = y, we get hx, A∗ xi = hA∗ x, xi = hx, A∗ xi, which im-
plies that hx, A∗ xi is real-valued.
Conversely, assume (2) is true. Since the mappings
(x, y) 7→ hy, A∗ xi
(x, y) 7→ hA∗ y, xi
are both sesquilinear forms on S × S (linear in the second variable, and conjugate
linear in the first variable), we apply the polarization identity:
1 3 kD E
hy, A∗ xi = ∑ i x + i k
y, A ∗
x + ik
y
4 k=0
1 3 D E
hA∗ y, xi = ∑ ik A∗ x + ik y , x + ik y
4 k=0
for all x, y ∈ D (A∗ ). Now, since A is Hermitian, the RHSs of the above equations
are equal; therefore, hy, A∗ xi = hA∗ y, xi, which is part (2).
Eqs (9.4)-(9.5) and lemma 9.1 set up a bijection between (closed) Hermitian
extensions of A and (closed) symmetric subspaces in D (A∗ ) D (A). Moreover, by
lemma 9.1, condition (9.6) is equivalent to
In particular,
k(A − λ ) xk2 ≥ |ℑ {λ }|2 kxk2 , ∀λ ∈ C. (9.9)
k(A − λ ) xk2
= h(A − a) x − ibx, (A − a) x − ibxi
= k(A − a) xk2 + |b|2 kxk2 − i (h(A − a) x, xi − hx, (A − a) xi)
= k(A − a) xk2 + |b|2 kxk2
≥ |b|2 kxk2 ;
Proof. Set B = A − λ ; then B is closed, and so is B−1 , i.e., the operator graphs G (B)
and G B−1 are closed in H ⊕ H . Therefore, ran (B) = dom(B−1 ) is closed in
k·kB−1 -norm. But by (9.9), B−1 is bounded on ran (B), thus the two norms k·k and
k·kB−1 are equivalent on ran (B). It follows that ran (B) is also closed in k·k-norm,
i.e., it is a closed subspace in H . The decomposition (9.10) follows from this.
D E
hy, (A − λ ) xi = A∗ − λ y, x , ∀x ∈ D (A) . (9.11)
232 9 Selfadjoint Extensions
hy − y0 , (A − λ ) xi = 0, ∀x ∈ D (A) .
ky − y0 k2 = hy − y0 , y − y0 i = 0.
Proof. Fix λ with ℑ {λ } > 0. For ℑ {λ } < 0, the argument is similar. We proceed to
verify that if η ∈ C, close enough to λ , then dim (ker (A∗ − η)) = dim (ker (A∗ − λ )).
The desired result then follows immediately.
Since A is closed, we have the following decomposition (by 9.1),
H = ran A − λ ⊕ ker (A∗ − λ ) . (9.13)
Now, pick x ∈ ker (A∗ − η), and suppose x ⊥ ker (A∗ − λ ); assuming kxk = 1. By
(9.13), ∃x0 ∈ D (A) s.t.
x = A − λ x0 . (9.14)
9.1 Extensions of Hermitian Operators 233
Then,
0 = h(A∗ − η) x, x0 i = hx, (A − η) x0 i
D E
= x, A − λ x0 − η − λ x0
= kxk2 − η − λ hx, x0 i
0 ≥ 1 − |η − λ | kx0 k2 ≥ 1 − |η − λ | |ℑ {λ }|−2
Similarly, we get the reversed inequality, and so dim (ker (A∗ − η)) = dim (ker (A∗ − λ )).
are called the deficiency spaces of A, and dimD± (A) are called the deficiency in-
dices.
234 9 Selfadjoint Extensions
Theorem 9.3 (von Neumann). Let A be a densely defined closed Hermitian opera-
tor acting in H . Then
where D(A∗ ) is identified with its graphG (A∗ ), thus a Hilbert space under the graph
inner product; and the decomposition in (9.17) refers to this Hilbert space.
and so D± (A), when identified with the graph of A∗ , are also closed sub-
D± (A∗ )
spaces in D (A∗ ).
Next, we verify the three subspaces on RHS of (9.17) are mutually orthogonal.
For all x ∈ D (A), and all x+ ∈ D+ (A) = ker (A∗ − i), we have
where the last step follows from x+ ⊥ ran (A + i) in H , see (9.10). Thus, D (A) ⊥
D+ (A) in D (A∗ ). Similarly, D (A) ⊥ D− (A) in D (A∗ ).
Moreover, if x+ ∈ D+ (A) and x− ∈ D− (A), then
But, by the decomposition H = ran (A + i) ⊕ ker (A∗ − i), eq. (9.10), there exist x0
and x+ satisfying (9.18). It remains to set x− := x − x0 − x+ , and to check x− ∈
D− (A). Indeed, by (9.18), we see that
A∗ x − Ax0 − i x+ = −i x + i x0 + i x+ ; i.e.,
A∗ (x − x0 − x+ ) = −i (x − x0 − x+ )
0 = hy, x+ i + hA∗ y, A∗ x+ i
= hy, x+ i + hA∗ y, i x+ i
= i (hi y, x+ i + hA∗ y, x+ i)
= i h(A∗ + i) y, x+ i , ∀x+ ∈ D+ (A) = ker (A∗ − i)
0 = hy, x− i + hA∗ y, A∗ x− i
= hy, x− i + hA∗ y, −i x− i
= −i (h−i y, x− i + hA∗ y, x− i)
= i h(A∗ − i) y, x− i , ∀x− ∈ D− (A) = ker (A∗ + i) ;
x1 + x2
y= ∈ D (A) .
2
236 9 Selfadjoint Extensions
x = x0 + x+ + x− ;
where x0 ∈ D (A), x+ ∈ ker (A∗ − z), and x− ∈ ker (A∗ − z). Then
and
(A∗ − z) x = (A − z) x0 + (z − z) x+ . (9.22)
Now, we start with (9.22). By the decomposition H = ran (A − z) ⊕ ker (A∗ − z),
there exist unique x0 and x+ such that (9.22) holds. This defines x0 and x+ . Then,
set
x− := x − x0 − x+ ;
A∗ (x − x0 − x+ ) = z (x − x0 − x+ )
Remark 9.2. In the general decomposition (9.21), if f = x+x+ +x− , g = y+y+ +y−
where f , g ∈ D (A), x+ , y+ ∈ ker (A∗ − z), and x− , y− ∈ ker (A∗ − z); then
hg, A∗ f i − hA∗ g, f i
= hy + y+ + y− , A∗ (x + x+ + x− )i − hA∗ (y + y+ + y− ) , x + x+ + x− i
= hy + y+ + y− , Ax + zx+ + zx− i − hAy + zy+ + zy− , x + x+ + x− i
9.1 Extensions of Hermitian Operators 237
Theorem 9.4 (von Neumann). Let A be a densely defined closed Hermitian opera-
tor in H .
(9.23)
Proof. By the discussion in (9.6) and (9.7), and lemma 9.1, it remains to character-
ize the closed symmetric subspaces S in D+ (A) ⊕ D− (A) (⊂ D (A∗ )). For this, let
x = x+ + x− , x± ∈ D± (A), then
Thus,
i.e., S is identified with the graph of a partial isometry, say U, with initial space in
D+ (A) and final space in D− (A).
Remark 9.3. M. Stone and von Neumann are the two pioneers who worked at the
same period. They were born at about the same time. Stone died at 1970’s and von
Neumann died in the 1950’s.
There is a simple criterion to test whether a Hermitian operator has equal deficiency
indices.
Theorem 9.5 (von Neumann). Let A be a densely defined closed Hermitian opera-
tor in H . Set d± = dim (D± (A)). Suppose AJ = JA, where J is a conjugation, then
d+ = d− . In particular, A has selfadjoint extensions.
Proof. Note that, by definition, we have hJx, yi = Jx, J 2 y = hJy, xi, for all x, y ∈
H.
We proceed to show that J commutes with A∗ . For this, let x ∈ D (A), y ∈ D (A∗ ),
then
hJA∗ y, xi = hJx, A∗ yi = hAJx, yi = hJAx, yi = hJy, Axi . (9.25)
9.2 Cayley Transform 239
It follows that x 7→ hJy, Axi is bounded, and Jy ∈ D (A∗ ). Thus, JD (A∗ ) ⊂ D (A∗ ).
Since J 2 = 1, D (A∗ ) = J 2 D (A∗ ) ⊂ JD (A∗ ); therefore, JD (A∗ ) = D (A∗ ). More-
over, (9.25) shows that JA∗ = A∗ J.
Now if x ∈ D+ (A), then
(A + i) x 7→ (A − i) x, ∀x ∈ D (A) (9.27)
is isometric. Equivalently,
CA x = (A − i) (A + i)−1 x (9.28)
Proof. By (9.3), ran (A ± i) are isometric to the graph of A, and the latter is closed
(as a subset in H ⊕ H ) since A is closed (i.e., G (A) is closed). Thus, ran (A ± i)
are closed in H . Note this is also a result of 9.1.
The mapping (9.28) being isometric follows from (9.26).
By (9.27), we have
which is (9.29).
CAe := CA ⊕U
U
eU ⊃ A.
is the Cayley transform of A
Given U as above, for all x ∈ D (A), x+ ∈ D+ (A), we have
Then,
= 2ix + (1 −U) x+
(1 +CAe ) ((A + i) x ⊕ x+ ) = ((A + i) x + x+ ) + ((A − i) x +Ux+ )
U
= 2Ax + (1 +U) x+ ;
and so
−1 1 1
i(1 +CAe )(1 −CAe ) x + (1 −U) x+ = Ax + (1 +U) x+ .
U U 2i 2
x = x0 + x+ + x−
y = y0 + y+ + y−
hy, A∗ xi − hA∗ y, xi
= hy0 + y+ + y− , Ax0 + i (x+ − x− )i − hAy0 + i (y+ − y− ) , x0 + x+ + x− i
= hy0 , Ax0 i − hAy0 , x0 i + hy0 , i (x+ − x− )i − hAy0 , x+ + x− i +
| {z } | {z }
0 0
hy+ + y− , Ax0 i − hi (y+ − y− ) , x0 i +
| {z }
0
hy+ + y− , i (x+ − x− )i − hi (y+ − y− ) , x+ + x− i
= 2i {hy+ , x+ i − hy− , x− i} . (9.30)
Hb = D+ (A)
ρ1 (x0 + x+ + x− ) = x+
ρ2 (x0 + x+ + x− ) = Ux+
∗
U =A
Af , where
D (Af
U)
∗
D Af
U = {x ∈ D (A ) : Uρ1 (x) = ρ2 (x)} .
Certain variations of theorem 9.8 are convenient in the boundary value problems
(BVP) of differential equations. In [DM91, GG91], a boundary triple (Hb , β1 , β2 )
is defined to satisfy
for all x, y ∈ D (A∗ ); and c0 is some nonzero constant. Also, see [JPT12c, JPT12b,
JPT12a].
The connection between (9.31) and (9.32) is via the bijection
β1 = ρ1 + ρ2
( )
ρ1 = β1 + iβ2 2
⇐⇒ . (9.33)
ρ2 = β1 − iβ2 β = ρ1 − ρ2
2
2i
244 9 Selfadjoint Extensions
hρ1 (x) , ρ1 (y)ib − hρ2 (x) , ρ2 (y)ib = 2i (hβ1 (x) , β2 (y)ib − hβ2 (x) , β1 (y)ib )
hρ1 , ρ1 ib − hρ2 , ρ2 ib
= hβ1 + iβ2 , β1 + iβ2 ib − hβ1 − iβ2 , β1 − iβ2 ib
= i hβ1 , β2 ib − i hβ2 , β1 ib + i hβ1 , β2 ib − i hβ2 , β1 ib
= 2i (hβ1 , β2 ib − hβ2 , β1 ib )
Theorem 9.9. Given a boundary triple (Hb , β1 , β2 ) satisfying (9.32), the family of
U ⊃ A is indexed by unitary operators U : Hb → Hb , such
selfadjoint extensions Af
that
∗
U =A
Af, where (9.34)
D (Af
U)
∗
D Af
U = {x ∈ D (A ) : (1 −U) β1 (x) = i (1 +U) β2 (x)} . (9.35)
d
Example 9.3. Let A = −i dx , and
D (A)
Then A∗ = −i dx
d
, where
D (A∗ )
D (A∗ ) = f : f , f 0 ∈ L2 (0, 1) .
i.e.,
fθ = −i d
A .
dx { f ∈D (A∗ ): f (0)=eiθ f (1)}
Example 9.4. Let A f = − f 00 , with D (A) = Cc∞ (0, ∞). Since A is Hermitian and A ≥
0, it follows that it has equal deficiency indices. Also, D (A∗ ) = f , f 00 ∈ L2 (0, ∞) ,
and A∗ f = − f 00 , ∀ f ∈ D (A∗ ).
For f , g ∈ D ∗ (A), we have
ˆ ∞ ˆ ∞
∗ 00
0 0 ∞
hg, A f i = − gf = − gf −g f 0 + g00 f
0 0
ˆ ∞
0 0
= g f (0) − g f (0) − g00 f
0
= g f 0 (0) − g0 f (0) + hA∗ g, f i
and so
hg, A∗ f i − hA∗ g, f i = g f 0 (0) − g0 f (0) .
where
1 + eiθ
z=i .
1 − eiθ
We take the convention that z = ∞ ⇐⇒ f 0 (0) = 0, i.e., the Neumann boundary
condition.
Example 9.5. A f = − f 00 , D (A) = Cc∞ (0, 1); then D (A∗ ) = f , f 00 ∈ L2 (0, 1) . In-
Thus,
where ! !
ϕ (0) ϕ 0 (0)
β1 (ϕ) = , β2 (ϕ) = .
ϕ 0 (1) ϕ (1)
∗
U =A
Af , where
D (Af
U)
∗
D Af
U = { f ∈ D (A ) : (1 −U) β1 ( f ) = i (1 +U) β2 ( f )} .
where
9.4 The Friedrichs Extension 247
! !
ϕ (0) ϕ 0 (0)
β1 (ϕ) = , β2 (ϕ) = .
ϕ (1) −ϕ 0 (1)
The selfadjoint boundary condition leads to
! !
f (0) f 0 (0)
(1 −U) = i (1 +U) .
f (1) − f 0 (1)
f 0 (0) = f 0 (1) = 0.
f (0) = f (1) = 0.
extends by continuity to
248 9 Selfadjoint Extensions
ϕe : H1 ,→ H
such that
kxk ≤ kxk1 , ∀x ∈ H1 . (9.38)
4. Define
e := A∗
A , where
dom(A)
e
e∗ ) ∩ H1 .
e := dom(A
dom(A)
e∗ , and L e = LA .
e=A
Then A A
Hence ϕ is continuous and the norm ordering passes to the completions of dom (A)
with respect to k·k1 and k·k. Therefore (9.38) holds.
Next, we verify that ϕe is injective (i.e., kerϕe = 0.) Suppose (xn ) ⊂ dom (A) s.t.
k·k k·k
1
xn −−→ x ∈ H1 , and xn −→ 0. We must show that kxk1 = 0. But
(3) Let (yn ) ⊂ dom (A), and kyn − yk1 → 0. For all x ∈ dom (A), we have
hy, xi1 = lim hyn , xi1 = lim hyn , Axi = hy, Axi .
n→∞ n→∞
9.4 The Friedrichs Extension 249
Equivalently,
Consequently,
= lim hAxm , yi
m→∞
= lim hxm , A∗ yi
m→∞
and
= lim hA∗ x, yn i
n→∞
= lim hAx,
e yn i = hAx,
e yi.
n→∞
Thus, A
e is Hermitian.
Fix y ∈ H . The map x 7−→ hy, xi, ∀x ∈ dom (A) ⊂ H1 , is linear and satisfies
hy, xi = hy , x 1
, ∀x ∈ H1 . (9.40)
250 9 Selfadjoint Extensions
In particular,
hy, xi = hy , x 1
= hy , Ax , ∀x ∈ dom (A) .
e xi = hy, xi , ∀y ∈ dom(A),
hAy, e ∀x ∈ dom (A) . (9.41)
1
e = H implies that A
Claim. ran(A) e is selfadjoint. In fact, for all x ∈ dom(A)
e and
e∗ ), we have
y ∈ dom(A
e∗ y = Ah,
where A e for some h ∈ dom(A), e = H . Thus,
e using the assumption ran(A)
hy − h, Axi
e = 0, ∀x ∈ dom(A);
e
hx, Axi
e = lim hx, Ax
e n i = lim hx, Axn i
n→∞ n→∞
Theorem 9.11. The Friedrichs extension of A is the unique selfadjoint operator sat-
isfying
e yi = hx, yi , ∀x ∈ dom(A),
hAx, e ∀y ∈ dom (A) .
1
See (9.41).
Fix x ∈ dom (A), and the above identify passes to y ∈ dom (C). Therefore, y ∈ B∗ =
B, and By = Cy. This shows C ⊂ B. Since
C = C ∗ ⊃ B∗ = B
h f , gi−1 = ξ f , ξg 1
, ∀ f , g ∈ H−1 . (9.46)
Proof. Since cx 7→ h·, cxi = c h·, xi, the mapping in (9.47) is linear.
For all x ∈ H0 , we have 8.3
(9.42)
|hy, xi0 | ≤ kxk0 kyk0 ≤ kxk0 kyk1 , ∀y ∈ H1 ; (9.48)
hence h·, xi0 is a bounded conjugate linear functional on H1 , i.e., h·, xi0 ∈ H−1 .
Moreover, by (9.48),
kh·, xi0 k−1 ≤ kxk0 . (9.49)
If h·, xi0 ≡ 0 in H−1 , then hy, xi0 = 0, for all y ∈ H1 . Since H1 is dense in H0 ,
it follows that x = 0 in H0 . Thus, (9.47) is injective.
Now, if f ⊥ {h·, xi0 : x ∈ H1 } in H−1 , then
(9.46) (9.44)
D E
h f , h·, xi0 i−1 = ξ f , ξh·,xi0 = ξf ,x 0
= 0, ∀x ∈ H1 . (9.50)
1
Thus, ξ f 0
= 0, since H1 is dense in H0 . Since id : H1 ,→ H0 is injective, and
ξ f ∈ H1 , it follows that ξ f 1 = 0. This, in turn, implies k f k−1 = 0, and so f = 0
in H−1 . Consequently, the image of H1 (resp. H0 as it contains H1 ) under (9.47)
is dense in H−1 .
Combining (9.42) and theorem 9.12, we get the triple of Hilbert spaces
(9.43) (9.47)
H1 ,−−−→ H0 ,−−−→ H−1 . (9.51)
1. All mappings in (9.51) are injective, continuous (in fact, contractive), hav-
ing dense images.
2. The map x 7→ ξh·,xi0 is a contraction from H1 ⊂ H0 into H0 . This follows
from the estimate:
(9.42)
ξh·,xi0 ≤ ξh·,xi0
0 1
(9.46)
= kh·, xi0 k−1
(9.49) (9.42)
≤ kxk0 ≤ kxk1 , ∀x ∈ H1 . (9.52)
f (x) = x, ξ f 1
, ∀x ∈ H1 , ∀ f ∈ H−1 . (9.53)
Combined with the order relation kxk−1 ≤ kxk0 for all x ∈ H0 , we see that
D E
hx, yi−1 = x, ξh·,yi0
0
≤ kxk0 ξh·,yi0
−1
= kxk0 kyk−1 .
≤ kxk0 kyk0 , , ∀x, y ∈ H0 .
Then,
In particular,
Proof.
hBy, Bxi1 = 0, ∀x ∈ H0 ;
equivalently,
hy, xi−1 = 0, ∀x ∈ H0 .
9.5 Rigged Hilbert Space 255
(9.56)
hx, Bxi0 = hBx, Bxi1 ≥ 0 =⇒ B ≥ 0.
(9.57) (9.49)
hx, Bxi0 = hBx, Bxi1 = hx, xi−1 ≤ hx, xi0 ;
Another argument:
1. A = A∗ , A ≥ 1.
2. dom (A) is dense in H1 and H0 , and ran (A) = H0 .
3. For all y ∈ dom (A), x ∈ H1 ,
In particular,
Proof. Part (1)-(3) are immediate by theorem 9.13. For (4), suppose A, B are selfad-
joint in H0 s.t.
(i) dom (A), dom (B) are contained in H1 , dense in H0 ;
256 9 Selfadjoint Extensions
(ii)
Thus, x 7→ hAx, yi0 is a bounded linear functional on dom (A), and so y ∈ dom (A∗ ) =
dom (A) and A∗ y = Ay = By; i.e., A ⊃ B. Since A, B are selfadjoint, then
B = B∗ ⊂ A∗ = A.
Therefore A = B.
D E
hx, yi1 = A1/2 x, A1/2 y , ∀x, y ∈ H1 . (9.60)
0
2. For all x, y ∈ H0 ,
D E
hx, yi−1 = A−1/2 x, A−1/2 y . (9.61)
0
kT xk ≤ kxk + kT xk ≤ (1 + c) kT xk
y ⊥ dom (A) in dom A1/2
m
D E
A1/2 y, A1/2 x = 0, ∀x ∈ dom A1/2
0
m
hy, Axi0 = 0, ∀x ∈ dom A1/2
m
y = 0 in H0
(9.58)
D E
hx, xi1 = hx, Axi0 = A1/2 x, A1/2 x , ∀x ∈ dom (A) .
0
258 9 Selfadjoint Extensions
Conclusion: (i) dom (A) is dense in H1 and dom A1/2 ; (ii) k·k1 and A1/2 · 0
agree on dom (A). Therefore the closures of dom (A) in H1 and dom A1/2 are
(2) Given x, y ∈ H0 ,
(9.57) (9.60)
D E
hx, yi−1 = A−1 x, A−1 y 1
= A−1/2 x, A−1/2 y .
0
(9.57) (9.60)
kAxk−1 = A−1 (Ax) 1
= kxk1 = A1/2 x .
0
D E
hy, xi1 = S1/2 y, S1/2 x , ∀x, y ∈ H1 . (9.64)
0
hy, Axi0 = hy, Sxi0 , ∀x ∈ dom (A) , ∀y ∈ dom S1/2 .
kxk21 ≥ LA kxk20 , ∀x ∈ H1 .
In particular,
hx, Sxi0 ≥ LA hx, xi0 , ∀x ∈ dom (S)
and so LS ≥ LA . Therefore, LA = LS .
For readers wishing to follow up sources, or to go in more depth with topics above,
we suggest: [AG93, Nel69, Dev72, DS88c, Kre46, Jor08, Rud73, Sto51b].
Chapter 10
Unbounded Graph-Laplacians
261
262 10 Unbounded Graph-Laplacians
variables and functions; or for efficient algorithms for computing them. By contrast,
for very large networks (like the Internet), variables are typically not known com-
pletely; – in most cases they may not even be well defined. In such applications,
data about them can only be collected by indirect means; hence random variables
and local sampling must be used as opposed to global processes.
Although such modern applications go far beyond the setting of large electrical
networks (even the case of infinite sets of vertices and edges), it is nonetheless true
that the framework of large electrical networks is helpful as a basis for the analysis
we develop below; and so our results will be phrased in the setting of large electrical
networks, even though the framework is much more general.
The applications of “large” or infinite graphs are extensive, counting just physics;
see for example [BCD06, RAKK05, KMRS05, BC05, TD03, VZ92].
1. (x, y) ∈ E ⇐⇒ (y, x) ∈ E; x, y ∈ V ;
2. # {y ∈ V | (x, y) ∈ E} is finite, and > 0 for all x ∈ V ;
3. (x, x) ∈
/ E; and
4. ∃ o ∈ V s.t. for all y ∈ V ∃ x0 , x1 , . . . , xn ∈ V with x0 = o, xn = y, (xi−1 , xi ) ∈
E, ∀i = 1, . . . , n. (This property is called connectedness.)
5. If a conductance function c is given we require cxi−1 xi > 0. See Definition
10.1 below.
1. c (e) ≥ 0, ∀e ∈ E; and
2. Given x ∈ V , cxy > 0, cxy = cyx , for all (xy) ∈ E.
If x ∈ V , we set
c (x) := ∑ cxy . (10.1)
(xy)∈E
0
0 −cx3 x2 c (x3 ) −cx3 x4 ··· ··· · · ·
0
.. . .. . .. . .. .. .. ..
.
0 . . . · · · (10.3)
. .. .. .. .. .. ..
..
. . . . . 0 . · · ·
.
0
.
. ··· 0 −cxn xn−1 c (xn ) −cxn xn+1 0 · · ·
. .. .. .. .. ..
.. . ··· ··· 0 . . . .
Remark 10.1 (Random Walk). If (V, E, c) is given as in Definition 10.2, then for
(x, y) ∈ E, set
cxy
pxy := (10.4)
c (x)
and note then pxy in (10.4) is a system of transition probabilities, i.e., ∑y pxy = 1,
∀x ∈ V , see Fig. 10.1 below.
y2
y3
p x y3
p x y2
x
p x y1
y1
1
hu, viHE := ∑∑ cxy (u (x) − u (y)) (v (x) − v (y)) (10.6)
2 (x,y)∈E
1
kuk2HE : = ∑∑ cxy |u (x) − u (y)|2
2 (x,y)∈E
(10.7)
Lemma 10.1. For all x, y ∈ V , there is a unique real-valued dipole vector vxy ∈ HE
s.t.
vxy , u HE
= u (x) − u (y) , ∀u ∈ HE .
Definition 10.3. Let H be a Hilbert space with inner product denoted h·, ·i, or
h·, ·iH when there is more than one possibility to consider. Let J be a countable
index set, and let w j j∈J be an indexed family of non-zero vectors in H . We say
that w j j∈J is a frame for H iff (Def.) there are two finite positive constants b1
A = AH : H −→ l 2 (J),
Au = w j, u H j∈J
(10.9)
u = A∗ Au = ∑ w j, u w j (10.11)
j∈J
Proof. The details are standard in the theory of frames; see the cited papers above.
Note that (10.8) for b1 = b2 = 1 simply states that A in (10.9) is isometric, and so
A∗ A = IH = the identity operator in H , and AA∗ = the projection onto the range
of A.
is a Parseval frame for the energy Hilbert space HE . For all u ∈ HE , we have the
following representation
Frames in HE consisting of our system (10.12) are not ONBs when resisters are
configured in non-linear systems of vertices, for example, resisters in parallel. See
Fig 10.2, and Example 10.1.
266 10 Unbounded Graph-Laplacians
V = Band V = Z2
Example 10.1. Let c01 , c02 , c12 be positive constants, and assign conductances on the
three edges (see Fig 10.3) in the triangle network.
c01 c02
c12
1 2
Fig. 10.3: The set vxy : (xy) ∈ E is not orthogonal.
√
In this case, wi j = ei j vi j , i < j, in the cyclic order is a Parseval frame but not
an ONB in HE [JT14b].
Note the corresponding Laplacian ∆ (= ∆c ) has the following matrix representa-
tion
c (0) −c01 −c02
M := −c01 c (1) −c12 (10.15)
∆ vxy = δx − δy .
Hence,
h itr
Mv01 = 1 −1 0
h itr
Mv02 = 1 0 −1
10.3 The Graph-Laplacian 267
h itr
Mv12 = 0 1 −1
Remark 10.2. The dipole vxy is unique in HE as an equivalence class, not a func-
tion on V . Note ker M = harmonic functions = constant (see (10.15)), and so
vxy + const = vxy in HE . Thus, the above frame vectors have non-unique repre-
sentations as functions on V .
Here we include some technical lemmas for graph Laplacian in the energy Hilbert
space HE .
Let G = (V, E, c) be as above; assume G is connected; i.e., there is a base point o
in V such that every x ∈ V is connected to o via a finite path of edges.
If x ∈ V , we set
1 if y = x
δx (y) = (10.16)
0 if y 6= x
vx := vx,o , ∀x ∈ V 0 .
Further, let
D2 := span δx x ∈ V , and
(10.17)
n o
DE := ∑x∈V 0 ξx vx finite support ; (10.18)
Moreover, we have
8.
c (x) = ∑t∼x cxt if y = x
δx , δy HE
= −cxy if (xy) ∈ E
if (xy) ∈
/ E, x 6= y
0
2
hϕ, ∆ ϕiHE = ∑ c2xy vxy , ϕ HE
.
(xy)∈E
Proof. Suppose ϕ = ∑ ϕxy vxy ∈ dom(∆ ). Note the edges are not oriented, and a
direct computation shows that
10.5 A 1D Example 269
2
hϕ, ∆ ϕiHE = 4 ∑ ϕxy .
x,y
Using the Parseval frames in Theorem 10.1, we have the following representation
1
ϕ= ∑ cxy vxy , ϕ v
HE xy
2
(xy)∈E | {z }
=:ϕxy
Note ϕ ∈ span vxy : x, y ∈ V , so the above equation contains a finite sum.
It follows that
2 2
hϕ, ∆ ϕiHE = 4 ∑ ϕxy = ∑ c2xy vxy , ϕ HE
(xy)∈E (xy)∈E
Proof. Follows from Lemma 10.4, and the characterization of Friedrichs extensions
of semibounded Hermitian operators (chapter 9); see, e.g., [DS88c, AG93, RS75].
10.5 A 1D Example
0o a1
/1o
a2
/2o
a3
/3 ··· no an+1
/ n+1 ···
x z y
0
−a2 a2 + a3 −a3
.
−a3 a3 + a4 . .
.. ..
. . −an (10.20)
−an an + an+1 −an+1
0 −an+1
..
.
..
.
.. ..
. .
.. ..
. .
That is,
(∆ u)0
= a1 (u0 − u1 )
(∆ u)n = an (un − un−1 ) + an+1 (un − un+1 ) (10.21)
= (an + an+1 ) un − an un−1 − an+1 un+1 , ∀n ∈ Z+ .
where ∞ 2
kuk2HE = ∑ an hvn−1,n , uiHE < ∞. (10.23)
n=1
√
forms a Parseval frame in HE . In
∞
Proof. By Theorem 10.1, the set an vn−1,n n=1
fact, the dipole vectors are
0 s ≤ n−1
vn−1,n (s) = ; n = 1, 2, . . . (10.24)
− 1 s≥n
an
√
forms an ONB in HE ; and u ∈ HE has the representation
∞
and so an vn−1,n n=1
∞
u= ∑ an hvn−1,n , uiHE vn−1,n
n=1
Below we compute the deficiency space in an example with index values (1, 1).
Lemma 10.6. Let (V, E, c = {an }) be as above. Let Q > 1 and set an := Qn , n ∈ Z+ ;
then ∆ has deficiency indices (1, 1).
and by induction,
1 1+Q 1
un+1 = + un − un−1 , n ∈ Z+
Qn+1 Q Q
Equivalently, as n → ∞, we have
1+Q 1 1 1
un+1 ∼ un − un−1 = 1 + un − un−1
Q Q Q Q
and so
1
un+1 − un ∼ (un − un−1 ) .
Q
Therefore, for the tail-summation, we have:
(Q − 1)2
∑ Qn (un+1 − un )2 = const ∑ Qn+2
<∞
n n
Next, we give a random walk interpretation of Lemma 10.6. See Remark 10.1,
and Fig 10.1.
1
is spanned by u = (un )∞
n=0 , un = Qn , n ∈ N; and of course k1/Qn k2HE < ∞.
Remark 10.4. For the domain of the Friedrichs extension ∆Fri , we have:
i.e.,
10.5 A 1D Example 273
( )
∞
2
dom(∆Fri ) = f ∈ HE | ∑ | f (x) − f (x + 1)| 2x
Q <∞ .
x=0
Proof. By Theorem 10.1, we have the following representation, valid for all f ∈
HE :
D x
E x
f = ∑ f , Q 2 v(x,x+1) Q 2 v(x,x+1)
x HE
and
h f , ∆ f iHE = ∑ | f (x) − f (x + 1)|2 Q2x .
x
The desired conclusion (10.25) now follows from Theorem 10.2. Also see e.g.
[DS88c, AG93].
Definition 10.5. Let G = (V, E, c) be a connected graph. The set of transition prob-
abilities (pxy ) is said to be reversible if there exists c : V → R+ s.t.
and then
cxy := c (x) pxy (10.27)
y y
cxy px y
x cxy ¢ x p x y¢
y¢ y¢
where
c (n) := an + an+1 (10.31)
and
an an+1
p− (n) := , p+ (n) := (10.32)
c (n) c (n)
are the left/right transition probabilities, as shown in Fig 10.6.
p- p+ p- p+
0 1 2 n-1 n n+1
Fig. 10.6: The transition probabilities p+ , p− , in the case of constant transition prob-
abilities, i.e., p+ (n) = p+ , and p− (n) = p− for all n ∈ Z+ .
Qn+1 Q
p+ := p+ (n) = = (10.34)
Qn + Qn+1 1+Q
10.5 A 1D Example 275
Qn 1
p− := p− (n) = n n+1
= (10.35)
Q +Q 1+Q
∆ = c (1 − P) . (10.37)
1 Q λ
fλ (n − 1) + fλ (n + 1) = 1 − n−1 f (n)
1+Q 1+Q Q (1 + Q) λ
and so
1+Q λ 1
fλ (n + 1) = − n fλ (n) − f (n − 1) . (10.38)
Q Q Q λ
This corresponds to the following matrix equation:
" # " #" #
1+Q
f (n + 1) Q − Qλn − Q1 f (n)
=
f (n) 1 0 f (n − 1)
" #" #
1+Q
Q − Q1 f (n)
∼ , as n → ∞.
1 0 f (n − 1)
That is, as n → ∞,
1+Q 1
fλ (n + 1) ∼ fλ (n) − f (n − 1) ;
Q Q λ
i.e.,
1
fλ (n + 1) ∼ f (n) ; (10.39)
Q λ
and so the tail summation of k fλ k2HE is finite. (See the proof of Lemma 10.6.) We
conclude that fλ ∈ HE .
Corollary 10.1. Let (V, E, ∆ ) be as in the lemma. The Friedrichs extension ∆Fri has
continuous spectrum [0, ∞).
and so every fλ , λ ∈ [0, ∞), is a generalized eigenfunction, i.e., the spectrum of ∆Fri
is purely continuous with Lebesgue measure, and multiplicity one.
The verification of (10.40) follows from eq. (10.38), i.e.,
1+Q λ 1
fλ (n + 1) = − n fλ (n) − f (n − 1) . (10.41)
Q Q Q λ
Set ˆ λ1
F[λ0 ,λ1 ] := fλ (·) dλ . (10.42)
λ0
For readers wishing to follow up sources, or to go in more depth with topics above,
we suggest: [Jor08, JP10, JP11a, JP11b, JP13a, JP13c, JT14b, RAKK05, Yos95].
Chapter 11
Reproducing Kernel Hilbert Space
A special family of Hilbert spaces H are reproducing kernel Hilbert spaces (RKHSs).
We say that H is a RKHS if H is a Hilbert space of functions on some set X such
that for every x in X, the linear mapping f 7−→ f (x) is continuous in the norm of
H.
We begin with a general
Let H be a RKHS, see Definition 11.1, hence S is a set, and H is a Hilbert space
of functions on S such that RS2 holds. Fix s ∈ S, and note that Es : H −→ C is
then a bounded linear operator, where C is a 1-dimensional Hilbert space. Hence its
adjoint Es∗ : C −→ H ∗ ' H is well-defined.
279
280 11 Reproducing Kernel Hilbert Space
The interest in positive definite functions has at least three roots: (i) Fourier analy-
sis, and harmonic analysis more generally, including the non-commutative variant
where we study unitary representations of groups; (ii) optimization and approxi-
mation problems, involving for example spline approximations as envisioned by I.
Schöenberg; and (iii) the study of stochastic (random) processes.
A stochastic process is an indexed family of random variables based on a fixed
probability space; in our present analysis, the processes will be indexed by some
group G; for example G = R, or G = Z correspond to processes indexed by real
time, respectively discrete time. A main tool in the analysis of stochastic processes
is an associated covariance function, see (11.1).
A process Xg g ∈ G is called Gaussian if each random variable Xg is Gaus-
sian, i.e., its distribution is Gaussian. For Gaussian processes we only need two
moments. So if we normalize, setting the mean equal to 0, then the process is deter-
mined by the covariance function. In general the covariance function is a function
on G × G, or on a subset, but if the process is stationary, the covariance function will
in fact be a positive definite function defined on G, or a subset of G. We will be using
three stochastic processes in the Memoir, Brownian motion, Brownian Bridge, and
the Ornstein-Uhlenbeck process, all Gaussian, or Ito integrals.
We outline a brief sketch of these facts below.
Let G be a locally compact group, and let (Ω , F , P) be a probability space, F a
sigma-algebra, and P a probability measure defined on F . A stochastic L2 -process is
a system of random variables Xg g∈G , Xg ∈ L2 (Ω , F , P). The covariance function
Proof. To stress the idea, we include the easy part of the theorem, and we refer to
[PS75] for the non-trivial direction:
Let λ1 , λ2 , . . . , λn ∈ C, and {gi }Ni=1 ⊂ G, then for all finite summations, we have:
2
N
g−1
∑ ∑ λi λ j cX i gj = E
∑ λi Xgi ≥ 0.
i j i=1
While each of the two extension problems has received a considerable amount of
attention in the literature, our emphasis here will be the interplay between the two
problems: Our aim is a duality theory; and, in the case G = Rn , and G = Tn =
Rn /Zn , we will state our theorems in the language of Fourier duality of abelian
groups: With the time frequency duality formulation of Fourier duality for G = Rn
we have that both the time domain and the frequency domain constitute a copy
of Rn . We then arrive at a setup such that our extension questions (i) are in time
domain, and extensions from (ii) are in frequency domain. Moreover we show that
each of the extensions from (i) has a variant in (ii). Specializing to n = 1, we arrive
of a spectral theoretic characterization of all skew-Hermitian operators with dense
domain in a separable Hilbert space, having deficiency-indices (1, 1).
11.3 The Reproducing Kernel Hilbert Space HF 283
xi−1 x j ≥ 0,
∑ ∑ ci c j F (11.5)
i j
For simplicity we focus on the case G = R, indicating the changes needed for gen-
eral Lie groups.
Definition 11.4. Fix 0 < a < ∞, set Ω := (0, a). Let F : Ω − Ω → C be a continuous
p.d. function. The reproducing kernel Hilbert space (RKHS), HF , is the completion
of
∑ c j F (· − x j ) : c j ∈ C (11.7)
finite
∑ ci F (· − xi ) , ∑ c j F (· − x j ) HF
= ∑ ∑ ci c j F (xi − x j ) , (11.8)
i j i j
Remark 11.1. Throughout, we use the convention that the inner product is conjugate
linear in the first variable, and linear in the second variable. When more than one
inner product is used, subscripts will make reference to the Hilbert space.
Notation. Inner product and norms will be denoted h·, ·i, and k·k respectively.
Often more than one inner product is involved, and subscripts are used for identifi-
cation.
Lemma 11.1. The reproducing kernel Hilbert space (RKHS), HF , is the Hilbert
completion of the functions
ˆ
Fϕ (x) = ϕ (y) F (x − y) dy, ∀ϕ ∈ Cc∞ (Ω ) , x ∈ Ω (11.9)
Ω
In particular,
ˆ ˆ
2
Fϕ HF
= ϕ (x)ϕ (y) F (x − y) dxdy, ∀ϕ ∈ Cc∞ (Ω ) (11.11)
Ω Ω
and ˆ
Fϕ , Fψ HF
= ϕ (x)Fψ (x) dx, ∀φ , ψ ∈ Cc∞ (Ω ). (11.12)
Ω
Proof. Apply standard approximation, see lemma 11.2 below.
Fix x ∈ (0, a), and set ϕn,x (t) := nϕ (n (t − x)). Then limn→∞ ϕn,x = δx , i.e., the
Dirac measure at x; and
Fϕn,x − F (· − x) HF
→ 0, as n → ∞. (11.13)
j n,x
0 x a
Recall, the following facts about HF , which follow from the general theory
[Aro50] of RKHS:
hF (· − x) , ξ iHF = ξ (x) , ∀ξ ∈ HF , ∀x ∈ Ω ,
Remark 11.2. It follows from the reproducing property that if Fφn → ξ in HF , then
Fφn converges uniformly to ξ in Ω . In fact
≤ kF (· − x)kHF Fφn − ξ HF
= F (0) Fφn − ξ HF
.
Lemma 11.3. Let F : (−a, a) → C be a continuous and p.d. function, and let HF
be the corresponding RKHS. Then:
´a
1. the integral Fϕ := 0 ϕ (y) F (· − y) dy is convergent in HF for all ϕ ∈
Cc (0, a); and
2. for all ξ ∈ HF , we have:
ˆ a
Fϕ , ξ HF
= ϕ (x)ξ (x) dx. (11.14)
0
Proof. For simplicity, we assume the following normalization F (0) = 1; then for
all y1 , y2 ∈ (0, 1), we have
Now, view the integral in (1) as a HF -vector valued integral. If ϕ ∈ Cc (0, a), this
´a
integral 0 ϕ (y) F (· − y) dy is the HF -norm convergent. Since HF is a RKHS,
h·, ξ iHF is continuous on HF , and it passes under the integral in (1). Using
d
Fϕ (x) = TF ϕ 0 (x) , ∀x ∈ (0, a) .
(11.18)
dx
´a
Proof. Since Fϕ (x) = 0 ϕ (y) F (x − y) dy, x ∈ (0, a); the desired assertion (11.18)
follows directly from the arguments in the proof of lemma 11.3.
for all finite system {ci } ⊂ C and {xi } ⊂ (0, a). Equivalently, for all ϕ ∈ Cc∞ (Ω ),
ˆ a 2 ˆ aˆ a
ϕ (y) ξ (y) dy ≤ A ϕ (x)ϕ (y) F (x − y) dxdy (11.20)
0 0 0
We will use these two conditions (11.19)(⇔(11.20)) when considering for example
the von Neumann deficiency-subspaces for skew Hermitian operators with dense
domain in HF .
2
Proof. Note, if ξ ∈ HF , the LHS in (11.20) is Fϕ , ξ HF
, and so (11.20) holds,
since h·, ξ iHF is continuous on HF .
If ξ is continuous on [0, a], and if (11.20) holds, then
ˆ a
HF 3 Fϕ 7−→ ϕ (y) ξ (y) dy
0
Lemma 11.4. The operator DF defines a skew-Hermitian operator with dense do-
main in HF .
Proof. By lemma 11.2, dom (DF ) is dense in HF . If ψ ∈ Cc∞ (0, a) and |t| <
dist (supp (ψ) , endpoints), then
ˆ aˆ a
2 2
Fψ(·+t) HF
= Fψ HF
= ψ (x)ψ (y) F (x − y) dxdy (11.21)
0 0
see (11.11), so
d 2
F HF
=0
dt ψ(·+t)
which is equivalent to
DF Fψ , Fψ HF
+ Fψ , DF Fψ HF
= 0. (11.22)
Lemma 11.5. Let F be a positive definite function on (−a, a), 0 < a < ∞ fixed. Let
DF be as in 11.5, so that DF ⊂ D∗F (lemma 11.4), where D∗F is the adjoint relative
to the HF inner product.
Then ξ ∈ HF (as a continuous function on [0, a]) is in dom (D∗F ) iff
2
ˆ aˆ a
2
DF Fϕ , ξ HF
≤ C Fϕ HF
=C ϕ (x)ϕ (y) F (x − y) dxdy (11.25)
0 0
2
ˆ a
(11.14)
DF Fϕ , ξ HF
= Fϕ 0 , ξ HF
= ϕ 0 (x)ξ (x) dx, ∀ϕ ∈ Cc∞ (0, a) (11.26)
0
So (11.25) holds ⇐⇒
ˆ a 2
2
ϕ 0 (x)ξ (x) dx ≤ C Fϕ HF
, ∀ϕ ∈ Cc∞ (0, a)
0
i.e.,
ˆ a 2
2
ϕ (x)ξ 0 (x) dx ≤ C Fϕ HF
, ∀ϕ ∈ Cc∞ (0, a) , and
0
ξ 0 as a distribution is in HF , and
ˆ a
ϕ (x)ξ 0 (x) dx = Fϕ , ξ 0 HF
0
0 ϕ (x)η (x) dx = Fϕ , η HF
, ∀ϕ ∈ Cc
∞ (0, a). See theorem 11.2.
∗
Corollary 11.2. h ∈ HF is in dom D2F iff h00 ∈ HF (h00 distribution derivative)
∗
and (D∗F )2 h = D2F h = h00 .
d 2
Proof. Application of (11.26) to DF Fϕ = Fϕ 0 , we have D2F Fϕ = Fϕ 00 = dx
Fϕ ,
∀ϕ ∈ Cc∞ (0, a), and
ˆ a
D2F Fϕ , h
HF
= Fϕ 00 , h H = ϕ 00 (x)h (x) dx
F
ˆ a 0
D ∗ E
= ϕ (x)h00 (x) dx = Fϕ , D2F h .
0 HF
Definition 11.6. [DS88c]Let D∗F be the adjoint of DF relative to HF inner product.
The deficiency spaces DEF ± consists of ξ± ∈ dom (D∗F ), such that D∗F ξ± = ±ξ± ,
i.e.,
n o
DEF ± = ξ± ∈ HF : Fψ 0 , ξ± HF
= Fψ , ±ξ± HF
, ∀ψ ∈ Cc∞ (Ω ) .
290 11 Reproducing Kernel Hilbert Space
and if
U (t) ξ − ξ
ξ ∈ dom (A) = ξ ∈ HF : s.t. lim exists
t→0 t
then
U (t) ξ − ξ
Aξ = s.t. lim . (11.28)
t→0 t
Now use Fx (·) = F (x − ·) defined in (0, a); and set
Corollary 11.4. Assume λ ∈ R is in the point spectrum of A, i.e., ∃ξλ ∈ dom (A),
ξλ 6= 0, s.t. Aξλ = iλ ξλ holds in HF , then ξλ = const · eλ , i.e.,
then since A ⊂ −D∗F , we get ξ ∈ dom (D∗F ) by lemma 11.5 and (11.27), and D∗F ξλ =
−ξλ0 where ξ 0 is the distribution derivative (see (11.24)); and by (11.27)
(11.31)
(Aξλ ) (x) = − (D∗F ξλ ) (x) = ξλ0 (x) = iλ ξλ (x) , ∀x ∈ (0, a) (11.32)
But by Schwartz, the distribution solutions to (11.33) are ξλ (x) = const · eλ (x) =
const · eiλ x .
In the considerations below, we shall be primarily concerned with the case when
a fixed continuous p.d. function F is defined on a finite interval (−a, a) ⊂ R. In this
case, by a Mercer operator, we mean an operator TF in L2 (0, a) where L2 (0, a) is
defined from Lebesgue measure on (0, a), given by
ˆ a
(TF ϕ) (x) := ϕ (y) F (x − y) dy, ∀ϕ ∈ L2 (0, a) , ∀x ∈ (0, a) . (11.34)
0
Lemma 11.6. Under the assumptions stated above, the Mercer operator TF is trace
class in L2 (0, a); and if F (0) = 1, then
Proof. This is an application of Mercer’s theorem [LP89, FR42, FM13] to the inte-
gral operator TF in (11.34). But we must check that F, on (−a, a), extends uniquely
by limit to a continuous p.d. function Fex on [−a, a], the closed interval. This is true,
and easy to verify, see e.g. [JPT14].
Corollary 11.5. Let F and (−a, a) be as in lemma 11.6. Then there is a sequence
(λn )n∈N , λn > 0, s.t. ∑n∈N λn = a, and a system of orthogonal functions {ξn } ⊂
L2 (0, a) ∩ HF such that
292 11 Reproducing Kernel Hilbert Space
ˆ a
ξn (x)ξm (x) dx = δn,m , n, m ∈ N. (11.37)
0
Fψ , Fϕ HF
= Fψ , TF−1 Fϕ 2
. (11.38)
Consequently,
−1/2
khkHF = kTF hk2 , ∀h ∈ HF . (11.39)
Proof. Note
Fψ , TF−1 Fϕ 2
= Fψ , TF−1 TF ϕ 2 = Fψ , ϕ 2
ˆ a ˆ a
= ψ (x) F (y − x) dx ϕ (y) dy
ˆ0 a ˆ a0
= ψ (x)ϕ (y) F (x − y) dxdy = Fψ , Fϕ HF
.
0 0
Proof. The functions ξn are in HF by theorem 11.2. We check directly (11.6) that
Dp p E p
λn ξn , λm ξm = λn λm ξn , T −1 ξm 2
HF
p
= λn λm λm−1 hξn , ξm i2 = δn,m .
Definition 11.7. Let G be a locally compact group, and let Ω be an open connected
subset of G. Let F : Ω −1 · Ω → C be a continuous positive definite function.
Consider a strongly continuous unitary representation U of G acting in some
Hilbert space K , containing the RKHS HF . We say that (U, K ) ∈ Ext (F) iff there
is a vector k0 ∈ K such that
i.e., Ext2 (F), consists of the solutions to problem (11.40) for which K %
HF , i.e., unitary representations realized in an enlargement Hilbert space.
(We write Fe ∈ HF for the vector satisfying hFe , ξ iHF = ξ (e), ∀ξ ∈ HF ,
where e is the neutral (unit) element in G, i.e., e g = g, ∀g ∈ G.)
2. In the special case, where G = Rn , and Ω ⊂ Rn is open and connected, we
consider
F : Ω −Ω → C
Remark 11.3. Note that (11.42) is consistent with (11.40): For if (U, K , k0 ) is a uni-
tary representation of G = Rn , such that (11.40) holds; then, by a theorem of Stone,
294 11 Reproducing Kernel Hilbert Space
Setting
dµ (λ ) := kPU (dλ ) k0 k2K , (11.44)
it is then immediate that we have: µ ∈ M+ (Rn ), and that the finite measure µ
satisfies
b (x) = F (x) , ∀x ∈ Ω − Ω .
µ (11.45)
Set n = 1: Start with a local p.d. continuous function F, and let HF be the corre-
sponding RKHS. Let Ext(F) be the compact convex set of probability measures on
R defining extensions of F.
We now divide Ext(F) into two parts, say Ext1 (F) and Ext2 (F).
All continuous p.d. extensions of F come from strongly continuous unitary rep-
resentations. So in the case of 1D, from unitary one-parameter groups of course, say
U(t).
Let Ext1 (F) be the subset of Ext(F) corresponding to extensions when the uni-
tary representation U(t) acts in HF (internal extensions), and Ext2 (F) denote the
part of Ext(F) associated to unitary representations U(t) acting in a proper enlarge-
ment Hilbert space K (if any), i.e., acting in a Hilbert space K corresponding to a
proper dilation of HF .
Our emphasis is von Neumann indices, and explicit formulas for partially defined
positive definite functions F, defined initially only on a symmetric interval (−a, a).
Among the cases of partially defined positive definite functions, the following ex-
ample F (x) = e−|x| , in the symmetric interval (−1, 1), will play a special role. The
present section is devoted to this example.
There are many reasons for this:
(i) It is of independent interest, and its type 1 extensions (see section §11.4)
can be written down explicitly.
(ii) Its applications include stochastic analysis [Itô06] as follows. Given a
random variable X in a process; if µ is its distribution, then there are two
11.5 The Case of e−|x| , |x| < 1 295
The notation “⊇” above refers to containment of operators, or rather of the respec-
tive graphs of the two operators; see [DS88c].
Lemma 11.7. Let F (x) = e−|x| , |x| < 1. Set Fx (y) := F (x − y), ∀x, y ∈ (0, 1); and
´1
Fϕ (x) = 0 ϕ (y) F (x − y) dy, ∀ϕ ∈ Cc∞ (0, 1). Define DF Fϕ = Fϕ 0 on the dense
subset
dom (DF ) = Fϕ : ϕ ∈ Cc∞ (0, 1) ⊂ HF .
(11.46)
moreover,
kξ+ kHF = kξ+ kHF = 1. (11.49)
Proof. (Note if Ω is any bounded, open and connected domain in Rn , then a locally
defined continuous p.d. function, F : Ω −Ω :→ C, extends uniquely to the boundary
∂ Ω := Ω \Ω by continuity [JPT14].)
In our current settings, Ω = (0, 1), and Fx (y) := F (x − y), ∀x, y ∈ (0, 1). Thus,
Fx (y) extends to all x, y ∈ [0, 1]. In particular,
are the two defect vectors, as shown in 11.3. Moreover, using the reproducing prop-
erty, we have
and (11.49) follows. For more details, see [JPT14, lemma 2.10.14].
Lemma 11.8. Let F be any continuous p.d. function on (−1, 1). Set
ˆ 1
h (x) = ϕ (y) F (x − y) dy, ∀ϕ ∈ Cc∞ (0, 1) ;
0
then
ˆ 1 ˆ 1
h (0) = ϕ (y) F (−y) dy, h (1) = ϕ (y) F (1 − y) dy (11.50)
0 0
ˆ 1 ˆ 1
h0 (0) = ϕ (y) F 0 (−y) dy, h0 (1) = ϕ (y) F 0 (1 − y) dy; (11.51)
0 0
then
ˆ 1 ˆ 1
h (0) = ϕ (y) e−y dy, h (1) = ϕ (y) ey−1 dy (11.52)
0 0
ˆ 1 ˆ 1
h0 (0) = ϕ (y) e−y dy, h0 (1) = − ϕ (y) ey−1 dy (11.53)
0 0
In particular,
/ HF .
however, δ0 , δ1 ∈
By von Neumann’s theory [DS88c] and lemma 11.5, the family of selfadjoint ex-
tensions of the Hermitian operator −iDF is characterized by
298 11 Reproducing Kernel Hilbert Space
Aθ h + c e−x + eiθ ex−1 = −i h0 + i c e−x − eiθ ex−1 , where
n o (11.56)
dom (Aθ ) := h + c e−x + eiθ ex−1 h ∈ dom (DF ) , c ∈ C .
Remark 11.5. In (11.56), h ∈ dom (DF ) (see (11.46)), and by 11.8, h satisfies the
boundary conditions (11.54)-(11.55). Also, by lemma 11.7, ξ+ = F0 = e−x , ξ− =
F1 = ex−1 , and kξ+ kHF = kξ− kHF = 1.
Proof. By assumption, eiλ x ∈ dom (Aθ ), so ∃hλ ∈ dom (DF ), and ∃cλ ∈ C s.t.
eiλ x = hλ (x) + cλ ex + eiθ ex−1 . (11.59)
1 − iλ 1−iλ
eiλ = eiθ = eiθ ei arg( 1+iλ ) (11.60)
1 + iλ
where
1 − iλ 2λ
arg = tan−1
1 + iλ λ2 −1
and (11.58) follows. For a discrete set of solutions, see figure 11.2.
-6 Π -4 Π -2 Π 2Π 4Π 6Π
-Π
hλ ∈ dom (DF ) , cλ ∈ C .
Remark 11.6. The corollary holds for all continuous p.d. functions F : (−a, a) → C.
Corollary 11.11. All selfadjoint extensions Aθ ⊃ −iDF have purely atomic spec-
trum; i.e.,
Λθ := spect (Aθ ) = discrete subset in R. (11.61)
i.e., all eigenvalues have multiplicity 1. (The set Λθ will be denoted {λn (θ )}n∈Z
following Fig. 11.2. )
For explicit computations regarding these points, see also Corollaries 11.16,
11.17, and 11.18 below.
ˆ 1
Fϕ (x) = ϕ (y) F (x − y) dy
0
ˆ 1
= ϕ (y) e−|x−y| dy, ∀ϕ ∈ Cc∞ (0, 1) .
0
Lemma 11.9. For all ϕ ∈ Cc∞ (0, 1), and all h, h00 ∈ HF , we have
= Fϕ , 21 h − h00 − 12 [W ]10
Fϕ , h HF 2
(11.63)
where " #
h Fϕ
W = det . (11.64)
h0 Fϕ 0
Setting l := Fϕ , we have
Proof. Note
ˆ 1
Fϕ , h HF
= ϕ (x) h (x) dx (reproducing property)
0
D E
d 2
= 12 I − dx Fϕ , h
2
11.5 The Case of e−|x| , |x| < 1 301
= Fϕ , 12 h − h00 − 21 [W ]10 .
2
Then
which is (11.65).
Corollary 11.13. eλ ∈ HF , ∀λ ∈ R.
But
ˆ 1 D E
1 d 2
ϕ (x) eλ (x) dx = 2 I− dx Fϕ , eλ
2
0
00 1
= Fϕ , 12 eλ − eλ − [W ]10
2 2
= 21 1 + λ 2 Fϕ , eλ − 12 −l (1) (1 + iλ ) eiλ − l (0) (1 − iλ ) ;
2
and we have
2
Fϕ , eλ 2
= |hTF ϕ, eλ i2 |2
D
1/2 1/2
E 2
= TF ϕ, TF eλ
2
1/2 2 1/2 2
≤ TF ϕ TF eλ (by Cauchy-Schwarz)
2 2
1/2 2
= hϕ, TF ϕi2 TF eλ
2
2 2 2
≤ Fϕ H keλ k2 = Fϕ H ;
F F
1/2 2
where we used the fact that TF eλ ≤ λ1 keλ k22 ≤ 1, since λ1 < 1 = the right
2
endpoint of the interval [0, 1] (see lemma 11.6), and keλ k2 = 1.
Therefore, the corollary follows.
Corollary 11.14. For all λ ∈ R, and all Fϕ , ϕ ∈ Cc∞ (0, 1), we have
1
1 + λ 2 Fϕ , eλ 2
Fϕ , eλ HF
= 2 (11.67)
+ 12 l (1) (1 + iλ ) eiλ + l (0) (1 − iλ ) .
= Fϕ , 12 eλ − e00λ − 12 [W ]10 .
Fϕ , eλ HF 2
where
1
eλ − e00λ = 1
1 + λ 2 eλ ; and
2 2
(11.65)
[W ]10 = −l (1) (1 + iλ ) eiλ − l (0) (1 − iλ ) , l := Fϕ .
Fϕ , eλ HF
= hϕ, eλ i2 . (11.68)
Proof. Eq. (11.68) follows from basic fact of the Mercer operator. See Lemma 11.6
and its corollaries. It suffices to note the following estimate:
ˆ 1
Fϕ00 (x) dx = F 0 ϕ (1) − Fϕ0 (0)
0
ˆ 1 ˆ 1
= −e−1 ey ϕ (y) dy − e−y ϕ (y) dy
0 0
= −Fϕ (1) − Fϕ (0) ≤ 2 Fϕ H
.
λ2 +3
heλ , eλ iHF = . (11.69)
2
1
1 + λ 2 Fϕ , eλ 2
Fϕ , eλ HF
=
2
1
+ l (1) (1 + iλ ) eiλ + l (0) (1 − iλ ) ; l := Fϕ . (11.70)
2
Fϕn , eλ HF
→ heλ , eλ iHF
1 1 −iλ
= 1+λ2 + e (1 + iλ ) eiλ + (1 − iλ )
2 2
1 λ2 +3
= 1+λ2 +1 =
.
2 2
304 11 Reproducing Kernel Hilbert Space
The approximation is justified since all the terms in the RHS of (11.70) satisfy
2
the estimate |· · · |2 ≤ C Fϕ H . See the proof of 11.13 for details.
F
Corollary 11.16. For all h ∈ HF , and all k ∈ dom TF−1 = Fϕ : ϕ ∈ Cc∞ (0, 1) ,
we have
1 1
hh, kiH = hh, ki0 + h0 , k0 0 + h (0)k (0) + h (1)k (1) (11.71)
2 2
1 1 λ2 +3
heλ , eλ iH = 1 + λ 2 + (1 + 1) =
2 2 2
as in (11.69).
1 − iλ iθ 1 − iµ iθ
eiλ = e , eiµ = e
1 + iλ 1 + iµ
and so
(1 − iµ) (1 + iλ )
ei(µ−λ ) = .
(1 + iµ) (1 − iλ )
Substitute this into (11.72) yields
−2 (1 + λ µ) 2 (1 + λ µ)
2 eλ , e µ H
= + = 0.
(1 + iµ) (1 − iλ ) (1 + iµ) (1 − iλ )
11.5 The Case of e−|x| , |x| < 1 305
Corollary 11.18. Let F (x) = e−|x| , |x| < 1. Let DF Fϕ = Fϕ 0 , ∀ϕ ∈ Cc∞ (0, 1), and
Then
2
Feθ (x) = ∑ e (x) , ∀x ∈ R (11.74)
λ ∈Λθ
λ2 +3 λ
is a continuous p.d. extension of F to the real line. Note that both sides in eq. (11.74)
depend on the choice of θ .
The type 1 extensions are indexed by θ ∈ [0, 2π) where Λθ is given in (11.73), see
also (11.58) in 11.9.
Corollary 11.19 (Sampling property of the set Λθ ). Let F (x) = e−|x| in |x| < 1,
HF , θ , and Λθ be as above. Let TF be the corresponding Mercer operator. Then for
all ϕ ∈ L2 (0, 1), we have
ϕb (λ ) iλ x
(TF ϕ) (x) = 2 ∑ e , for all x ∈ (0, 1) .
λ ∈Λθ
λ2 +3
Remark 11.7. Note that the system {eλ | λ ∈ Λθ } is orthogonal in HF , but not in
L2 (0, 1).
Proof. We saw that Aθ has pure atomic spectrum. By (11.15), the set
(r )
2
e : λ ∈ Λθ
λ +3 λ
2
1
Fθ (x) = ∑ heλ , FiHF eλ (x)
λ ∈Λθ keλ k2HF
2
= ∑ e (x) , ∀x ∈ [0, 1] . (11.75)
λ ∈Λθ
λ2 +3 λ
where heλ , FiHF = eλ (0) = 1 by the reproducing property. But the RHS of (11.75)
extends to R. See figure 11.3.
306 11 Reproducing Kernel Hilbert Space
1.0
0.8
0.6
0.4
0.2
-4 -2 2 4
θ is an ONB in HF .
2
[0, 2π), and let Λθ be as above; then e
λ 2 +3 λ
| λ ∈ Λ
For readers wishing to follow up sources, or to go in more depth with topics above,
we suggest: [AD86, JPT14, Nus75, Rud63].
Appendix A
An overview of Functional Analysis books (cast
of characters)
Below we offer a list of related Functional Analysis books; they cover a host of
diverse areas of functional analysis and applications, some different from what we
select here: Our comments are telegraphic-review form:
307
308 A An overview of Functional Analysis books (cast of characters)
This book is two books bound as one; and in the lovely format from Dover. Part
1: metric spaces, and normed linear spaces. Part 2: Lebesgue integration and ba-
sic functional analysis. Numerous examples are sprinkled through the text. To get
the most out of this book, it helps if you have already seen many of the results
presented elsewhere. History: The book came from original notes from Andrei Kol-
mogorov’s lectures given at Moscow’s Lomonosov University in the 1940’s, and it
still stands as timely introduction to real and functional analysis. Strengths: step by
step presentation of all the key concepts needed in the subject; proceeding all the
way from set theory to Fredholm integral equations. Offers a wonderful and refresh-
ing insight. Contents (sample): Elements of Set Theory; Metric and Topological
Spaces; Normed and Topological Linear Spaces; Linear Functionals and Linear Op-
erators; Elements of Differential Calculus in Linear Spaces; Measure, Measurable
Functions, Integral; Indefinite Lebesgue Integral, Differentiation Theory; Spaces of
Summable Functions; Trigonometric Series, Fourier Transformation; Linear Inte-
gral Equations.
to the present. That is in the book!! And it offers an upbeat outlook for the future.
It has been tested in the class room, – it is really user-friendly. At the end of each
chapter P Lax offers personal recollections; – little known stories of how several of
the pioneers in the subject have been victims, – in the 30ties and the 40ties, of Nazi
atrocities. The writing is crisp and engaged, – the exercises are great; – just right for
students to learn from. This is the book to teach from.
equations which again starts with some examples of problems from the 19th century
mathematicians. The presentation of Fredholm’s method is a gem.
analysis, to approximation theory, and to special functions. The last two chapters
illustrate the theory with a systematic study of (infinite × infinite) Jacobi matrices;
i.e., tri-diagonal infinite matrices; assumed formally selfadjoint (i.e., Hermitian).
Sample results: A dichotomy: Their von Neumann indices must be (0, 0) or (1, 1).
Some of the first known criteria for when they are one or the other are given; plus a
number of applications to classical analysis.
In chapters 4 and 8 above we have cited pioneers in quantum physics, the founda-
tions of quantum mechanics. The most central here are Heisenberg (matrix mechan-
ics), Schrödinger (wave mechanics, the Schrödinger equation), and Dirac (Dirac’s
equation is a relativistic wave equation, describes all spin-½ massive particles free
form, as well as electromagnetic interactions). We further sketched von Neumann’s
discovery of the equivalence of the answers given by Heisenberg and Schrödinger,
and the Stone-von Neumann uniqueness theorem. The relevant papers and books
are as follows: [Hei69, Sch32a, vN31, HN28, Dir35, Dir47].
A An overview of Functional Analysis books (cast of characters) 315
Functional Analysis
O
u *
Linear Mathematical
Operators (A) Physics (C)
O c ; O
Harmonic Representation
Analysis (B) i 5 Theory (D)
) # { u
Probability Theory / Statistics (E)
Fig. A.1
(A)
bounded differential operators, ODE/PDE
x
unbounded generators of diffusion
y
geometry
x
spectral theory Schrödinger operators
y
spectral representation wave operators
single operators scattering operators
system of operators
operator commutation relations
Table A.1
316 A An overview of Functional Analysis books (cast of characters)
(B)
analysis / synthesis
Fourier analysis, wavelet analysis
commutative
x
non-commutative Applications:
y
signal processing
physics
statistics
analysis on fractals
Table A.2
(C)
quantum physics
x
classical mechanics
y
quantum information
statistical physics states and decomposition
x
quantum field theory equilibrium: Gibbs, KMS, ...
y
relativistic
x
non-relativistic
y
Table A.3
(D)
Table A.4
A An overview of Functional Analysis books (cast of characters) 317
(E)
discrete
continuous
Gaussian
Brownian motion
stochastic processes non-Gaussian
Lévy
solutions of diffusion equations with the use of
functional integrals (i.e., probability measure on
infinite-dimensional spaces such as C (R) or
Schwartz space S)
Table A.5
Appendix B
Often cited above
mathematical ideas originate in empirics. But, once they are
conceived, the subject begins to live a peculiar life of its own
and is . . . governed by almost entirely aesthetical motivations.
In other words, at a great distance from its empirical source, or
after much “abstract” inbreeding, a mathematical subject is in
danger of degeneration. Whenever this stage is reached the only
remedy seems to me to be the rejuvenating return to the source:
the reinjection of more or less directly empirical ideas.
– von Neumann
Inside the book, the following authors are cited frequently, W. Arveson, S. Banach,
P. Dirac, K. Friedrichs, I. Gelfand, W. Heisenberg, D. Hilbert, K. Ito, M. Krein,
E. Nelson, R. Phillips, E. Schrödinger, H.A. Schwarz, L. Schwartz, J. Schwartz, I.
Segal, M. Stone, J. von Neumann, N. Wiener. Below a short bio.
W. Arveson. [Arv72, Arv98]. Cited in connection with C∗ -algebras and their states
and representations.
William Arveson (1934 – 2011); known for his work on completely positive maps,
and their extensions; powerful generalizations of the ideas of Banach, Krein, and
Stinespring . An early results in this area is an extension theorem for completely
positive maps with values in the algebra of all bounded operators. This theorem
led to injectivity of von-Neumann algebras in general, and work by Alain Connes
relating injectivity to hyperfiniteness. In a series of papers in the 60’s and 70’s,
Arveson introduced non-commutative analogues of several concepts from classical
harmonic analysis including the Shilov and. Choquet boundaries.
319
320 B Often cited above
Stefan Banach (1892 – 1945), one of the founders of modern functional analysis
and one of the original members of the Lwów School of Mathematics, in Poland be-
tween the two World Wars. His 1932 book, Théorie des opérations linéaires (The-
ory of Linear Operations), is the first monograph on the general theory of functional
analysis.
P. Dirac. [Dir35, Dir47]. Cited in connection with the “Dirac equation” and
especially our notation for vectors and operators in Hilbert space, as well as the
axioms of observables, states and measurements.
Paul Adrien Maurice Dirac (1902 – 1984), an English theoretical physicist; fun-
damental contributions to the early development of both quantum mechanics and
quantum electrodynamics. He was the Lucasian Professor of Mathematics at the
University of Cambridge. Notable discoveries, the Dirac equation, which describes
the behavior of fermions and predicted the existence of antimatter. Dirac shared the
Nobel Prize in Physics for 1933 with Erwin Schrödinger, “for the discovery of new
productive forms of atomic theory.”
Werner Karl Heisenberg (1901 – 1976); one of the key creators of quantum me-
chanics. A 1925 paper was a breakthrough. In the subsequent series of papers with
Max Born and Pascual Jordan, this matrix formulation of quantum mechanics took a
mathematical rigorous formulation. In 1927 he published his uncertainty principle.
Heisenberg was awarded the Nobel Prize in Physics for 1932 “for the creation of
quantum mechanics.” He made important contributions to the theories of the hydro-
dynamics of turbulent flows, the atomic nucleus.
322 B Often cited above
D. Hilbert. [Hil24, Hil22, Hil02]. Cited in connection with the early formulations
of the theory of operators in (what is now called) Hilbert space. The name Hilbert
space was suggested by von Neumann who studied with Hilbert in the early
1930ties, before he moved to the USA. (The early papers by von Neumann are in
German.)
David Hilbert (1862 – 1943) is recognized as one of the most influential and univer-
sal mathematicians of the 19th and early 20th centuries. Discovered and developed
invariant theory and axiomatization of geometry. In his his 1900 presentation of a
collection of research problems, he set the course for much of the mathematical
research of the 20th century.
K. Ito. [Itô07, Itô04]. Cited in connection with Brownian motion, Ito-calculus, and
stochastic processes. Making connection to functional analysis via the theory of
semigroups of operators (Hille and Phillips.)
Erwin Rudolf Josef Alexander Schrödinger (1887 – 1961), a Nobel Prize in physics.
– quantum theory forming the basis of wave mechanics: he formulated the wave
equation (stationary and time-dependent Schrödinger equation) , and he Schrödinger
proposed an original interpretation of the physical meaning of the wave function;
formalized the notion of entanglement. He was critical the conventional Copen-
hagen interpretation of quantum mechanics (using e.g. the paradox of Schrödinger’s
cat).
B Often cited above 323
J. Schwartz. [DS88b, DS88c, DS88a]. Is the Schwartz of the book set “linear
operators” by Dunford and Schwartz. Vol II [DS88c] is one of the best presentation
of the theory of unbounded operators.
AD86. Daniel Alpay and Harry Dym, On applications of reproducing kernel spaces to the
Schur algorithm and rational J unitary factorization, I. Schur methods in operator
theory and signal processing, Oper. Theory Adv. Appl., vol. 18, Birkhäuser, Basel,
1986, pp. 89–159. MR 902603 (89g:46051)
AG93. N. I. Akhiezer and I. M. Glazman, Theory of linear operators in Hilbert space, Dover
Publications Inc., New York, 1993, Translated from the Russian and with a preface by
Merlynd Nestell, Reprint of the 1961 and 1963 translations, Two volumes bound as
one. MR 1255973 (94i:47001)
AH13. Giles Auchmuty and Qi Han, Spectral representations of solutions of linear elliptic
equations on exterior regions, J. Math. Anal. Appl. 398 (2013), no. 1, 1–10. MR
2984310
AJ12. Daniel Alpay and Palle E. T. Jorgensen, Stochastic processes induced by singular op-
erators, Numer. Funct. Anal. Optim. 33 (2012), no. 7-9, 708–735. MR 2966130
AJL13. Daniel Alpay, Palle Jorgensen, and Izchak Lewkowicz, Extending wavelet filters: infi-
nite dimensions, the nonrational case, and indefinite inner product spaces, Excursions
in harmonic analysis. Volume 2, Appl. Numer. Harmon. Anal., Birkhäuser/Springer,
New York, 2013, pp. 69–111. MR 3050315
AJLM13. Daniel Alpay, Palle Jorgensen, Izchak Lewkowicz, and Itzik Marziano, Representation
formulas for Hardy space functions through the Cuntz relations and new interpola-
tion problems, Multiscale signal analysis and modeling, Springer, New York, 2013,
pp. 161–182. MR 3024468
AJS14. Daniel Alpay, Palle Jorgensen, and Guy Salomon, On free stochastic processes and
their derivatives, Stochastic Process. Appl. 124 (2014), no. 10, 3392–3411. MR
3231624
AJSV13. Daniel Alpay, Palle Jorgensen, Ron Seager, and Dan Volok, On discrete analytic func-
tions: products, rational functions and reproducing kernels, J. Appl. Math. Comput.
41 (2013), no. 1-2, 393–426. MR 3017129
Alp92. Daniel Alpay, On linear combinations of positive functions, associated reproducing
kernel spaces and a non-Hermitian Schur algorithm, Arch. Math. (Basel) 58 (1992),
no. 2, 174–182. MR 1143167 (92m:46039)
325
326 B Often cited above
Cas13. Peter G. Casazza, The Kadison-Singer and Paulsen problems in finite frame theory,
Finite frames, Appl. Numer. Harmon. Anal., Birkhäuser/Springer, New York, 2013,
pp. 381–413. MR 2964016
Cas14. Peter G. Casazza, Consequences of the marcus/spielman/stivastava solution to the
kadison-singer problem, arXiv:1407.4768 (2014).
CH08. D. A. Croydon and B. M. Hambly, Local limit theorems for sequences of simple
random walks on graphs, Potential Anal. 29 (2008), no. 4, 351–389. MR 2453564
(2010c:60218)
CJK+ 12. Jianxin Chen, Zhengfeng Ji, David Kribs, Zhaohui Wei, and Bei Zeng, Ground-state
spaces of frustration-free Hamiltonians, J. Math. Phys. 53 (2012), no. 10, 102201, 15.
MR 3050570
CM13. B. Currey and A. Mayeli, The Orthonormal Dilation Property for Abstract Parseval
Wavelet Frames, Canad. Math. Bull. 56 (2013), no. 4, 729–736. MR 3121682
Con90. John B. Conway, A course in functional analysis, second ed., Graduate Texts in Math-
ematics, vol. 96, Springer-Verlag, New York, 1990. MR 1070713 (91e:46001)
CRKS79. P. Cotta-Ramusino, W. Krüger, and R. Schrader, Quantum scattering by external met-
rics and Yang-Mills potentials, Ann. Inst. H. Poincaré Sect. A (N.S.) 31 (1979), no. 1,
43–71. MR 557051 (81h:81129)
CW14. K. L. Chung and R. J. Williams, Introduction to stochastic integration, second ed.,
Modern Birkhäuser Classics, Birkhäuser/Springer, New York, 2014. MR 3136102
dB68. Louis de Branges, Hilbert spaces of entire functions, Prentice-Hall Inc., Englewood
Cliffs, N.J., 1968. MR 0229011 (37 #4590)
dBR66. Louis de Branges and James Rovnyak, Canonical models in quantum scattering theory,
Perturbation Theory and its Applications in Quantum Mechanics (Proc. Adv. Sem.
Math. Res. Center, U.S. Army, Theoret. Chem. Inst., Univ. of Wisconsin, Madison,
Wis., 1965), Wiley, New York, 1966, pp. 295–392. MR 0244795 (39 #6109)
Dev59. Allen Devinatz, On the extensions of positive definite functions, Acta Math. 102 (1959),
109–134. MR 0109992 (22 #875)
Dev72. , The deficiency index of a certain class of ordinary self-adjoint differential
operators, Advances in Math. 8 (1972), 434–473. MR 0298102 (45 #7154)
DHL09. Dorin Dutkay, Deguang Han, and David Larson, A duality principle for groups, J.
Funct. Anal. 257 (2009), no. 4, 1133–1143. MR 2535465 (2010i:22006)
Dir35. P. A. M. Dirac, The electron wave equation in de-Sitter space, Ann. of Math. (2) 36
(1935), no. 3, 657–669. MR 1503243
Dir47. , The Principles of Quantum Mechanics, Oxford, at the Clarendon Press, 1947,
3d ed. MR 0023198 (9,319d)
Dix81. Jacques Dixmier, Von neumann algebras, North Holland, 1 1981.
DM85. H. Dym and H.P. McKean, Fourier series and integrals, Probability and Mathematical
Statistics, ACADEMIC PressINC, 1985.
DM91. V. A. Derkach and M. M. Malamud, Generalized resolvents and the boundary value
problems for Hermitian operators with gaps, J. Funct. Anal. 95 (1991), no. 1, 1–95.
MR 1087947 (93d:47046)
dO09. César R. de Oliveira, Intermediate spectral theory and quantum dynamics, Progress in
Mathematical Physics, vol. 54, Birkhäuser Verlag, Basel, 2009. MR 2723496
328 B Often cited above
DS88a. N. Dunford and J.T. Schwartz, Linear operators, spectral operators, Linear Operators,
Wiley, 1988.
DS88b. Nelson Dunford and Jacob T. Schwartz, Linear operators. Part I, Wiley Classics Li-
brary, John Wiley & Sons Inc., New York, 1988, General theory, With the assistance
of William G. Bade and Robert G. Bartle, Reprint of the 1958 original, A Wiley-
Interscience Publication. MR 1009162 (90g:47001a)
DS88c. , Linear operators. Part II, Wiley Classics Library, John Wiley & Sons Inc.,
New York, 1988, Spectral theory. Selfadjoint operators in Hilbert space, With the as-
sistance of William G. Bade and Robert G. Bartle, Reprint of the 1963 original, A
Wiley-Interscience Publication. MR 1009163 (90g:47001b)
Emc00. G.G. Emch, Mathematical and conceptual foundations of 20th-century physics, North-
Holland Mathematics Studies, Elsevier Science, 2000.
EO13. Martin Ehler and Kasso A. Okoudjou, Probabilistic frames: an overview, Finite
frames, Appl. Numer. Harmon. Anal., Birkhäuser/Springer, New York, 2013, pp. 415–
436. MR 2964017
Fan10. Mark Fannes, An introduction to quantum probability, Theoretical foundations of
quantum information processing and communication, Lecture Notes in Phys., vol. 787,
Springer, Berlin, 2010, pp. 1–38. MR 2762151 (2012d:81174)
Fef71. Charles Fefferman, Characterizations of bounded mean oscillation, Bull. Amer. Math.
Soc. 77 (1971), 587–588. MR 0280994 (43 #6713)
FL28. Kurt Friedrichs and Hans Lewy, Über die Eindeutigkeit und das Abhängigkeitsgebiet
der Lösungen beim Anfangswertproblem linearer hyperbolischer Differentialgleichun-
gen, Math. Ann. 98 (1928), no. 1, 192–204. MR 1512400
FM13. J. C. Ferreira and V. A. Menegatto, Positive definiteness, reproducing kernel Hilbert
spaces and beyond, Ann. Funct. Anal. 4 (2013), no. 1, 64–88. MR 3004212
FR42. W. H. J. Fuchs and W. W. Rogosinski, A note on Mercer’s theorem, J. London Math.
Soc. 17 (1942), 204–210. MR 0008270 (4,272g)
Fri80. Kurt Otto Friedrichs, Spectral theory of operators in Hilbert space, Applied Mathe-
matical Sciences, vol. 9, Springer-Verlag, New York-Berlin, 1980, Corrected reprint.
MR 635783 (82j:47001)
Fug74. Bent Fuglede, Boundary minimum principles in potential theory, Math. Ann. 210
(1974), 213–226. MR 0357827 (50 #10293b)
Geo09. H. Georgi, Weak interactions and modern particle theory, Dover Books on Physics
Series, Dover Publications, Incorporated, 2009.
GG59. I. M. Gel0 fand and M. I. Graev, Geometry of homogeneous spaces, representations of
groups in homogeneous spaces and related questions of integral geometry. I, Trudy
Moskov. Mat. Obšč. 8 (1959), 321–390; addendum 9 (1959), 562. MR 0126719 (23
#A4013)
GG91. V. I. Gorbachuk and M. L. Gorbachuk, Boundary value problems for operator differ-
ential equations, Mathematics and its Applications (Soviet Series), vol. 48, Kluwer
Academic Publishers Group, Dordrecht, 1991, Translated and revised from the 1984
Russian original. MR 1154792 (92m:34133)
B Often cited above 329
GG02. M. Gadella and F. Gómez, A unified mathematical formalism for the Dirac formula-
tion of quantum mechanics, Found. Phys. 32 (2002), no. 6, 815–869. MR 1917012
(2003j:81070)
GJ60. I. M. Gel0 fand and A. M. Jaglom, Integration in functional spaces and its applications
in quantum physics, J. Mathematical Phys. 1 (1960), 48–69. MR 0112604 (22 #3455)
Gli60. James G. Glimm, On a certain class of operator algebras, Trans. Amer. Math. Soc. 95
(1960), 318–340. MR 0112057 (22 #2915)
Gli61. James Glimm, Type I C∗ -algebras, Ann. of Math. (2) 73 (1961), 572–612. MR 0124756
(23 #A2066)
GLS12. Martin J. Gander, Sébastien Loisel, and Daniel B. Szyld, An optimal block iterative
method and preconditioner for banded matrices with applications to PDEs on irregular
domains, SIAM J. Matrix Anal. Appl. 33 (2012), no. 2, 653–680. MR 2970224
GS60. I. M. Gel0 fand and Do-šin Sya, On positive definite distributions, Uspehi Mat. Nauk
15 (1960), no. 1 (91), 185–190. MR 0111991 (22 #2849)
HdSS12. Seppo Hassi, Hendrik S. V. de Snoo, and Franciszek Hugon Szafraniec (eds.), Operator
methods for boundary value problems, London Mathematical Society Lecture Note
Series, vol. 404, Cambridge University Press, Cambridge, 2012. MR 3075434
Hei69. W Heisenberg, Über quantentheoretische umdeutung kinematischer und mechanischer
beziehungen.(1925) in: G, Ludwig, Wellenmechanik, Einführung und Originaltexte,
Akademie-Verlag, Berlin (1969), 195.
Hid80. Takeyuki Hida, Brownian motion, Applications of Mathematics, vol. 11, Springer-
Verlag, New York, 1980, Translated from the Japanese by the author and T. P. Speed.
MR 562914 (81a:60089)
Hil02. David Hilbert, Mathematical problems, Bull. Amer. Math. Soc. 8 (1902), no. 10, 437–
479. MR 1557926
Hil22. , Die logischen Grundlagen der Mathematik, Math. Ann. 88 (1922), no. 1-2,
151–165. MR 1512123
Hil24. , Die Grundlagen der Physik, Math. Ann. 92 (1924), no. 1-2, 1–32. MR
1512197
HJL+ 13. Deguang Han, Wu Jing, David Larson, Pengtong Li, and Ram N. Mohapatra, Dilation
of dual frame pairs in Hilbert C∗ -modules, Results Math. 63 (2013), no. 1-2, 241–250.
MR 3009685
HKLW07. Deguang Han, Keri Kornelson, David Larson, and Eric Weber, Frames for undergrad-
uates, Student Mathematical Library, vol. 40, American Mathematical Society, Provi-
dence, RI, 2007. MR 2367342 (2010e:42044)
HN28. D. Hilbert and J. v. Neumann, Über die Grundlagen der Quantenmechanik, Math. Ann.
98 (1928), no. 1, 1–30. MR 1512390
Itô04. Kiyosi Itô, Stochastic processes, Springer-Verlag, Berlin, 2004, Lectures given at
Aarhus University, Reprint of the 1969 original, Edited and with a foreword by Ole
E. Barndorff-Nielsen and Ken-iti Sato. MR 2053326 (2005e:60002)
Itô06. Kiyoshi Itô, Essentials of stochastic processes, Translations of Mathematical Mono-
graphs, vol. 231, American Mathematical Society, Providence, RI, 2006, Translated
from the 1957 Japanese original by Yuji Ito. MR 2239081 (2007i:60001)
330 B Often cited above
Itô07. Kiyosi Itô, Memoirs of my research on stochastic analysis, Stochastic analysis and
applications, Abel Symp., vol. 2, Springer, Berlin, 2007, pp. 1–5. MR 2397781
JÓ00. Palle E. T. Jorgensen and Gestur Ólafsson, Unitary representations and Osterwalder-
Schrader duality, The mathematical legacy of Harish-Chandra (Baltimore, MD, 1998),
Proc. Sympos. Pure Math., vol. 68, Amer. Math. Soc., Providence, RI, 2000, pp. 333–
401. MR 1767902 (2001f:22036)
Jor08. Palle E. T. Jorgensen, Essential self-adjointness of the graph-Laplacian, J. Math. Phys.
49 (2008), no. 7, 073510, 33. MR 2432048 (2009k:47099)
Jør14. Palle E. T. Jørgensen, A universal envelope for Gaussian processes and their kernels,
J. Appl. Math. Comput. 44 (2014), no. 1-2, 1–38. MR 3147727
JP10. Palle E. T. Jorgensen and Erin Peter James Pearse, A Hilbert space approach to effec-
tive resistance metric, Complex Anal. Oper. Theory 4 (2010), no. 4, 975–1013. MR
2735315 (2011j:05338)
JP11a. Palle E. T. Jorgensen and Erin P. J. Pearse, Resistance boundaries of infinite networks,
Random walks, boundaries and spectra, Progr. Probab., vol. 64, Birkhäuser/Springer
Basel AG, Basel, 2011, pp. 111–142. MR 3051696
JP11b. , Spectral reciprocity and matrix representations of unbounded operators, J.
Funct. Anal. 261 (2011), no. 3, 749–776. MR 2799579
JP12. P. E. T. Jorgensen and A. M. Paolucci, q-frames and Bessel functions, Numer. Funct.
Anal. Optim. 33 (2012), no. 7-9, 1063–1069. MR 2966144
JP13a. Palle E. T. Jorgensen and Erin P. J. Pearse, A discrete Gauss-Green identity for un-
bounded Laplace operators, and the transience of random walks, Israel J. Math. 196
(2013), no. 1, 113–160. MR 3096586
JP13b. , A discrete Gauss-Green identity for unbounded Laplace operators, and the
transience of random walks, Israel J. Math. 196 (2013), no. 1, 113–160. MR 3096586
JP13c. , Multiplication operators on the energy space, J. Operator Theory 69 (2013),
no. 1, 135–159. MR 3029492
JPT12a. Palle Jorgensen, Steen Pedersen, and Feng Tian, Restrictions and extensions of semi-
bounded operators, Complex Analysis and Operator Theory (2012) (English).
JPT12b. , Translation representations and scattering by two intervals, J. Math. Phys. 53
(2012), no. 5, 053505, 49. MR 2964262
JPT12c. Palle E.T. Jorgensen, Steen Pedersen, and Feng Tian, Momentum operators in two in-
tervals: Spectra and phase transition, Complex Analysis and Operator Theory (2012)
(English).
JPT14. Palle Jorgensen, Steen Pedersen, and Feng Tian, Harmonic analysis of a class of re-
producing kernel Hilbert spaces arising from groups, arXiv:1401.4782 (2014).
JT14a. Palle Jorgensen and Feng Tian, Frames and factorization of graph laplacians,
arXiv:1404.1424 (2014).
JT14b. , Frames and factorization of graph laplacians, arXiv:1404.1424 (2014).
JT14c. , Infinite networks and variation of conductance functions in discrete lapla-
cians, arXiv:1404.4686 (2014).
Kat95. Tosio Kato, Perturbation theory for linear operators, Classics in Mathematics,
Springer-Verlag, Berlin, 1995, Reprint of the 1980 edition. MR 1335452 (96a:47025)
B Often cited above 331
1982), Lectures in Appl. Math., vol. 21, Amer. Math. Soc., Providence, RI, 1985,
pp. 219–253. MR 789292 (86j:81051)
Mac88. , Induced representations and the applications of harmonic analysis, Harmonic
analysis (Luxembourg, 1987), Lecture Notes in Math., vol. 1359, Springer, Berlin,
1988, pp. 16–51. MR 974302 (90c:22021)
Mac92. , The scope and history of commutative and noncommutative harmonic analy-
sis, History of Mathematics, vol. 5, American Mathematical Society, Providence, RI;
London Mathematical Society, London, 1992. MR 1171011 (93g:22006)
Mac09. Barbara D. MacCluer, Elementary functional analysis, Graduate Texts in Mathematics,
vol. 253, Springer, New York, 2009. MR 2462971 (2010b:46001)
MSS13. Adam Marcus, Daniel A Spielman, and Nikhil Srivastava, Interlacing families ii:
Mixed characteristic polynomials and the kadison-singer problem, arXiv:1306.3969
(2013).
Nel59a. Edward Nelson, Analytic vectors, Ann. of Math. (2) 70 (1959), 572–615. MR 0107176
(21 #5901)
Nel59b. , Regular probability measures on function space, Ann. of Math. (2) 69 (1959),
630–643. MR 0105743 (21 #4479)
Nel67. , Dynamical theories of Brownian motion, Princeton University Press, Prince-
ton, N.J., 1967. MR 0214150 (35 #5001)
Nel69. , Topics in dynamics. I: Flows, Mathematical Notes, Princeton University
Press, Princeton, N.J., 1969. MR 0282379 (43 #8091)
Nus75. A. Edward Nussbaum, Extension of positive definite functions and representation of
functions in terms of spherical functions in symmetric spaces of noncompact type of
rank 1, Math. Ann. 215 (1975), 97–116. MR 0385473 (52 #6334)
OH13. Anatol Odzijewicz and Maciej Horowski, Positive kernels and quantization, J. Geom.
Phys. 63 (2013), 80–98. MR 2996399
OR07. V. S. Olkhovsky and E. Recami, Time as a quantum observable, Internat. J. Modern
Phys. A 22 (2007), no. 28, 5063–5087. MR 2371443 (2009b:81008)
Ørs79. Bent Ørsted, Induced representations and a new proof of the imprimitivity theorem, J.
Funct. Anal. 31 (1979), no. 3, 355–359. MR 531137 (80d:22007)
Par82. K.R. Parthasarathy, Probability measures on metric spaces, AMS Chelsea Publishing
Series, Acad. Press, 1982.
Par09. K. R. Parthasarathy, An invitation to quantum information theory, Perspectives in math-
ematical sciences. I, Stat. Sci. Interdiscip. Res., vol. 7, World Sci. Publ., Hackensack,
NJ, 2009, pp. 225–245. MR 2581746 (2011d:81061)
Phe01. Robert R. Phelps, Lectures on choquet’s theorem (lecture notes in mathematics), 2nd
ed., Springer, 5 2001.
Pow75. Robert T. Powers, Simplicity of the C∗ -algebra associated with the free group on two
generators, Duke Math. J. 42 (1975), 151–156. MR 0374334 (51 #10534)
Pre89. P. M. Prenter, Splines and variational methods, Wiley Classics Library, John Wiley
& Sons, Inc., New York, 1989, Reprint of the 1975 original, A Wiley-Interscience
Publication. MR 1013116 (90j:65001)
PS75. K. R. Parthasarathy and K. Schmidt, Stable positive definite functions, Trans. Amer.
Math. Soc. 203 (1975), 161–174. MR 0370681 (51 #6907)
B Often cited above 333
RAKK05. G. J. Rodgers, K. Austin, B. Kahng, and D. Kim, Eigenvalue spectra of complex net-
works, J. Phys. A 38 (2005), no. 43, 9431–9437. MR 2187996 (2006j:05186)
RS75. M. Reed and B. Simon, Methods of modern mathematical physics: Fourier analysis,
self-adjointness, Fourier analysis, Self-adjointness, no. v. 2, Academic Press, 1975.
RSN90. Frigyes Riesz and Béla Sz.-Nagy, Functional analysis, Dover Books on Advanced
Mathematics, Dover Publications, Inc., New York, 1990, Translated from the sec-
ond French edition by Leo F. Boron, Reprint of the 1955 original. MR 1068530
(91g:00002)
Rud63. Walter Rudin, The extension problem for positive-definite functions, Illinois J. Math. 7
(1963), 532–539. MR 0151796 (27 #1779)
Rud70. , An extension theorem for positive-definite functions, Duke Math. J. 37 (1970),
49–53. MR 0254514 (40 #7722)
Rud73. , Functional analysis, McGraw-Hill Book Co., New York, 1973, McGraw-Hill
Series in Higher Mathematics. MR MR0365062 (51 #1315)
Rud87. , Real and complex analysis, third ed., McGraw-Hill Book Co., New York,
1987. MR 924157 (88k:00002)
Rud90. , Fourier analysis on groups, Wiley Classics Library, John Wiley & Sons Inc.,
New York, 1990, Reprint of the 1962 original, A Wiley-Interscience Publication. MR
1038803 (91b:43002)
Sak71. Shôichirô Sakai, C∗ -algebras and W ∗ -algebras, Springer-Verlag, New York, 1971,
Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 60. MR MR0442701 (56
#1082)
Sch32a. E. Schrödinger, Sur la théorie relativiste de l’électron et l’interprétation de la mé-
canique quantique, Ann. Inst. H. Poincaré 2 (1932), no. 4, 269–310. MR 1508000
Sch32b. , Sur la théorie relativiste de l’électron et l’interprétation de la mécanique
quantique, Ann. Inst. H. Poincaré 2 (1932), no. 4, 269–310. MR 1508000
Sch40. , A method of determining quantum-mechanical eigenvalues and eigenfunc-
tions, Proc. Roy. Irish Acad. Sect. A. 46 (1940), 9–16. MR 0001666 (1,277d)
Sch57. Laurent Schwartz, Théorie des distributions à valeurs vectorielles. I, Ann. Inst. Fourier,
Grenoble 7 (1957), 1–141. MR 0107812 (21 #6534)
Sch58. , La fonction aléatoire du mouvement brownien, Séminaire Bourbaki; 10e an-
née: 1957/1958. Textes des conférences; Exposés 152 à 168; 2e éd. corrigée, Exposé
161, vol. 23, Secrétariat mathématique, Paris, 1958. MR 0107312 (21 #6037)
Sch70. H. A. Schwarz, Gesammelte mathematische abhandlungen (ams chelsea publishing),
2nd ed., American Mathematical Society, 1 1970.
Sch95. Laurent Schwartz, Les travaux de L. Gårding sur les équations aux dérivées partielles
elliptiques, Séminaire Bourbaki, Vol. 2, Soc. Math. France, Paris, 1995, pp. Exp. No.
67, 175–182. MR 1609224
Sch99. E. Schrödinger, About Heisenberg uncertainty relation (original annotation by A. An-
gelow and M.-C. Batoni), Bulgar. J. Phys. 26 (1999), no. 5-6, 193–203 (2000), Transla-
tion of Proc. Prussian Acad. Sci. Phys. Math. Sect. 19 (1930), 296–303. MR 1782215
(2001i:81002)
SD13. F. A. Shah and Lokenath Debnath, Tight wavelet frames on local fields, Analysis
(Berlin) 33 (2013), no. 3, 293–307. MR 3118429
334 B Often cited above
337
338 Index