Pub Operator-Theory
Pub Operator-Theory
These lecture notes are based on the courses Operator Theory developed at
King’s College London by G. Barbatis, E.B. Davies and J.A. Erdos, and
Functional Analysis II developed at the University of Sussex by P.J. Bushell,
D.E. Edmunds and D.G. Vassiliev. As usual, all errors are entirely my re-
sponsibility. The same applies to the accompanying exercise sheets.
1 Introduction
Spaces
By IF we will always denote either the field IR of real numbers or the field
C of complex numbers.
1.3. Examples
1. Any finite-dimensional normed space is a Banach space.
2. Let lp , 1 ≤ p ≤ ∞, denote the vector space of all sequences x =
(xk )∞
k=1 , xk ∈ IF, such that
∞
!1/p
p
X
kxkp := |xk | < ∞, 1 ≤ p < ∞,
k=1
1
Then k · kp is a norm on lp and lp , 1 ≤ p ≤ ∞, are Banach spaces.
3. Let C([0, 1]) be the space of all continuous functions on [0, 1] and
Z 1 1/p
p
kf kp := |f (t)| dt , 1 ≤ p < ∞,
0
kf k∞ := sup |f (t)|.
0≤t≤1
So, (fn ) is a Cauchy sequence with respect to k·kp . Now, suppose there exists f ∈ C([0, 1])
such that kf − fn kp → 0. Taking into account that for an arbitrary δ ∈]0, 1/2[ there exists
N such that
fn = 1 on [1/2 + δ, 1], ∀n > N,
we obtain
Z 1 Z 1
p
|1 − f (t)| dt = |fn (t) − f (t)|p dt ≤ kf − fn kpp → 0 as n → ∞.
1/2+δ 1/2+δ
Since the first integral is independent of n it has to be zero, which implies that f = 1 on
[1/2 + δ, 1] and hence on ]1/2, 1] since δ was arbitrary. A similar argument shows that
f = 0 on [0, 1/2[. Thus f is discontinuous at t = 1/2 and we obtain a contradiction.
2
1.5. Theorem For any normed space (X, k · k) there exists a Banach
space (X 0 , k · k0 ) and a linear isometry from X onto a dense linear subspace
of X 0 . Two Banach spaces in which (X, k · k) can be so imbedded are iso-
morphic .
Operators
1.7. Definition Let X and Y be vector spaces. A map B : X → Y is
called a linear operator (map) if
B(λx + µz) = λBx + µBz, ∀x, z ∈ X, ∀λ, µ ∈ IF.
1.8. Theorem Let X and Y be normed spaces. For a linear operator
B : X → Y the following statements are equivalent:
(i) B is continuous;
(ii) B is continuous at 0;
(iii) there exists a constant C < +∞ such that kBxk ≤ Ckxk, ∀x ∈ X.
3
(prove these relations!).
B(X) := B(X, X)
A vector space E is called an algebra if for any pair (x, y) ∈ E × E a unique product
xy ∈ E is defined with the properties
(xy)z = x(yz),
x(y + z) = xy + xz,
(x + y)z = xz + yz,
λ(xy) = (λx)y = x(λy),
for all x, y, z ∈ E and scalars λ.
E is called an algebra with identity if it contains an element e such that for all x ∈ E,
ex = xe = x, ∀x ∈ E.
4
1.12. Theorem Let X and Y be Banach spaces and D be a dense lin-
ear subspace of X. Let A be a bounded linear operator from D (equipped
with the X−norm) into Y. Then there exists a unique extension of A to
a bounded linear operator Ā : X → Y defined on the whole X; moreover
kAk = kĀk.
1.13. Example Let X = Y be the space L2 ([a, b]), i.e. the completion
of D = (C([a, b]), k · k2 ), (−∞ < a < b < +∞). Consider the operator
A : C([a, b]) → C([a, b]) ⊂ L2 ([a, b]),
Z b
(Af )(t) = k(t, τ )f (τ )dτ, where k ∈ C([a, b]2 ).
a
Theorem 1.12 allows one to extend this operator to a bounded linear operator
Ā : L2 ([a, b]) → L2 ([a, b]), since D is dense in its completion X = L2 ([a, b]).
We only need to prove that A : D → D is a bounded operator.
Denote
K = max 2 |k(t, τ )|.
(t,τ )∈[a,b]
5
We say that the series converges absolutely if
∞
X
kxn k < ∞.
n=1
Ker(A) := {x ∈ X : Ax = 0}.
6
2.2. Theorem (Banach) Let X and Y be Banach spaces and let B ∈
B(X, Y ) be one-to-one and onto (i.e. Ker(B) = {0} and Ran(B) = Y ).
Then the inverse operator B −1 : Y → X is bounded, i.e. B −1 ∈ B(Y, X).
Proof: The proof of this fundamental result can be found in any textbook on
functional analysis. 2
Proof: (i)
B −1 T −1 T B = B −1 B = IX , T BB −1 T −1 = T T −1 = IZ .
(ii) According to (i) we have to prove only that the invertibility of T B implies
the invertibility of B and T . Let S : X → X be the inverse of T B, i.e.
ST B = T BS = I. It is clear that ST is a left inverse of B. Since B and
T commute, we have BT S = I, i.e. T S is a right inverse of B. Hence B
is invertible and its inverse equals ST = T S. Similarly we prove that T is
invertible.
(iii)
B −1 T = B −1 T BB −1 = B −1 BT B −1 = T B −1 . 2
7
2.4. Lemma Let X be a Banach space, B ∈ B(X) and kBk < 1. Then
I − B is invertible,
∞
(I − B)−1 = Bn
X
(2.1)
n=0
and k(I − B)−1 k ≤ 1/(1 − kBk).
and
kA−1 k
k(A + B)−1 k ≤
1 − kBkkA−1 k
Proof: We have
A + B = A(I + A−1 B) = (I + BA−1 )A
and kA−1 Bk ≤ kA−1 kkBk < 1, kBA−1 k ≤ kBkkA−1 k < 1. Now it is left to
apply Lemmas 2.3(i) and 2.4. 2
8
The spectrum
In this subsection we will deal only with complex vector spaces, i.e. with
the case IF = C, if it is not stated otherwise.
σ(B) = C\ρ(B),
2.8. Examples
1. If the space X is finite-dimensional then the spectrum of B ∈ B(X) co-
incides with the set of all its eigenvalues, i.e. with the set of zeros of the
determinant of a matrix corresponding to B − λI (with respect to a given
basis of X).
2. Let X be one of the Banach spaces C([0, 1]), Lp ([0, 1]), 1 ≤ p ≤ ∞, and
let B be defined by the formula
9
2.9. Lemma Let X be a Banach space and B ∈ B(X). Then σ(B) is
a compact set and
σ(B) ⊂ {λ ∈ C : |λ| ≤ kBk}. (2.2)
Proof: Suppose |λ| > kBk. Then
Proof:
10
commute with each other.
Proof: It is clear that the operators A, B, A − µI and B − λI commute. Now
the desired result follows from Lemma 2.3(iii). 2
For any normed space X we denote by X ∗ its dual (= adjoint) space, i.e.
the space of all bounded linear functionals on X, i.e. B(X, IF).
i.e.
1
(F (λ) − F (λ0 )) − F 0 (λ0 )
→ 0 as λ → λ0 ;
λ − λ0
(ii) any λ0 ∈ Ω has a neighbourhood where F (λ) can be represented by an
absolutely convergent series
∞
(λ − λ0 )n Fn (λ0 ), Fn (λ0 ) ∈ Z;
X
F (λ) =
n=0
Ω 3 λ 7→ G(F (λ)) ∈ C
is holomorphic (= analytic) in Ω.
If Z = B(X, Y ) for some Banach spaces X and Y , then for the operator–
valued function F the above statements are equivalent to
(iv) for any x ∈ X and g ∈ Y ∗ the complex–valued function
Ω 3 λ 7→ g(F (λ)x) ∈ C
11
is holomorphic (= analytic) in Ω.
Proof: It is clear that each of the statements (i) and (ii) implies (iii). (Why?)
It is also obvious that (iii) implies (iv). Indeed, for any x ∈ X and g ∈ Y ∗
the mapping
B(X, Y ) 3 B 7→ g(Bx) ∈ C
is a bounded linear functional on B(X, Y ). The opposite implications are
very non–trivial. We will not give the proof here. It can be found, e.g., in E.
Hille & R.S. Phillips, Functional Analysis and Semi–Groups, Ch. III, Sect.
2. 2
2.16. Remark We will not use the non–trivial part of Theorem 2.14. We
will only use the fact that each of the statements (i) and (ii) implies (iii) and
(iv). In the proof of the following theorem we will check that the resolvent
satisfies both (i) and (ii).
Proof: Let us take an arbitrary λ0 ∈ ρ(B). It follows from Lemma 2.5 and the
proof of Lemma 2.9 that the function R(B; ·) is bounded in a neighbourhood
of λ0 . From the resolvent equation (Lemma 2.11) we obtain
R(B; λ) → R(B; λ0 ) as λ → λ0
12
and
R(B; λ) − R(B; λ0 )
lim = lim R(B; λ)R(B; λ0 ) = R2 (B; λ0 ).
λ→λ0 λ − λ0 λ→λ0
Thus (2.4) is valid for any point λ0 ∈ ρ(B), i.e R(B; ·) is analytic in ρ(B).
Note that (2.3) and Lemma 2.5 imply the following Taylor expansion
∞
(λ − λ0 )n Rn+1 (B; λ0 ), if |λ − λ0 | < 1/k(B − λ0 I)−1 k. (2.7)
X
R(B; λ) =
n=0
Further,
∞
X
−1 −1 −n n
k − λR(B; λ) − Ik = k(I − λ B) − Ik =
λ B
≤
n=1
∞
|λ|−1 kBk kBk
|λ|−n kBkn =
X
−1
= → 0, as |λ| → ∞
n=1 1 − |λ| kBk |λ| − kBk
2.18. Lemma Let X be a normed space. Then for any z ∈ X there exists
g ∈ X ∗ such that g(z) = kzk and kgk = 1.
13
Proof: Let us suppose that σ(B) = ∅ and take arbitrary x ∈ X\{0}, g ∈ X ∗ .
It follows from Theorem 2.17 that the function
Combining Lemmas 2.9, 2.19 and Theorem 2.17 we obtain the following re-
sult.
2.21. Remark Let (ak ) be a sequence of real numbers. We will use the
following notation:
lim
n→∞
inf an := n→∞
lim inf ak , lim sup an := n→∞
lim sup ak .
k≥n n→∞ k≥n
2.22. Definition Let B ∈ B(X). The spectral radius r(B) of the operator
B is defined by the equality
14
This is the radius of the minimal disk centred at 0 and containing σ(B).
Since B−λI is not invertible and the operators in the RHS commute, B n −λn I
is not invertible (see Lemma 2.3(ii)). So, λ ∈ σ(B) implies λn ∈ σ(B n ).
Consequently
and therefore
r(B) ≤ lim inf kB n k1/n .
n→∞
15
Since R(B; ·) is analytic in C\σ(B) (see Theorem 2.17), so is f (see Theorem
2.14 and Remark 2.16). Hence, (2.11) is valid if |λ| > r(B). Taking λ =
aeiθ , a > r(B) and integrating the series for λm+1 f (λ) term by term with
respect to θ we have
Z 2π Z ∞
2π X
am+1 ei(m+1)θ f (aeiθ )dθ = − am−n ei(m−n)θ g(B n x)dθ =
0 0 n=0
∞ Z 2π
am−n g(B n x) ei(m−n)θ dθ = −2πg(B m x).
X
−
n=0 0
Thus
1 Z 2π m+1
|g(B m x)| ≤ a |g(R(B; aeiθ )x)|dθ ≤ am+1 M (a)kgkkxk,
2π 0
where
M (a) := sup kR(B; aeiθ )k
0≤θ≤2π
(see (1.1)). Taking g ∈ X ∗ such that g(B m x) = kB m xk, kgk = 1 (see Lemma
2.18) we obtain
kB m xk ≤ am+1 M (a)kxk, ∀x ∈ X,
i.e.
kB m k ≤ am+1 M (a), ∀m ∈ IN.
Consequently
lim sup kB m k1/m ≤ a, if a > r(B),
m→∞
16
|λ| < rf }, where rf > 0 is some given number. Then we have the following
Taylor expansion
∞
an λn , |λ| < rf ,
X
f (λ) =
n=0
We need to check that this definition makes sense. Let us take ε > 0 such
that r(B) + ε < rf . It follows from the spectral radius formula that for suffi-
ciently large n we have kB n k < (r(B)+ε)n . Therefore the power series (2.12)
is absolutely convergent. Hence, f (B) ∈ B(X) (see Theorems 1.10 and 1.15).
17
The RHS is well defined and belongs to B(X) if ∂Ω σ(B) = ∅. Here and
T
below the integrals are understood as norm limits of the corresponding Rie-
mann sums. This motivates the following definition.
2.27. Proposition The RHS of (2.14) does not depend on the choice of
an admissible set satisfying (2.13).
Proof: Let Ω and Ω1 be admissible sets satisfying (2.13). Then there ex-
ists an admissible set Ω0 such that
\
σ(B) ⊂ Ω0 ⊂ Cl(Ω0 ) ⊂ Ω Ω1 ⊂ ∆f .
18
Thus,
1 Z 1 Z
g f (λ)R(B; λ)xdλ − f (λ)R(B; λ)xdλ = 0, ∀g ∈ X ∗ .
2πi ∂Ω 2πi ∂Ω0
Consequently
1 Z 1 Z
f (λ)R(B; λ)xdλ = f (λ)R(B; λ)xdλ, ∀x ∈ X,
2πi ∂Ω 2πi ∂Ω0
(see Lemma 2.18), i.e. (2.15) holds. 2
Our aim now is to show that Definition 2.26 does not contradict the “com-
mon sense”, i.e. that in the situations where we now what f (B) is, (2.14)
gives the same result.
m 1 Z
(B − λ0 I) = − (λ − λ0 )m R(B; λ)dλ, ∀m ∈ Z. (2.17)
2πi ∂Ω
In particular
1 Z
R(B; λ0 ) = (B − λ0 I)−1 = − (λ − λ0 )−1 R(B; λ)dλ.
2πi ∂Ω
Proof: Let us denote the RHS of (2.17) by Am . Since λ0 6∈ Cl(Ω), the function
f (λ) := (λ − λ0 )m is analytic in some neighbourhood of Cl(Ω) and
1 Z
(λ − λ0 )m dλ = 0.
2πi ∂Ω
Using this equality and the resolvent equation (Lemma 2.11)
we obtain
1 Z
Am = −R(B; λ0 ) (λ − λ0 )m dλ −
2πi ∂Ω
1 Z
R(B; λ0 ) (λ − λ0 )m+1 R(B; λ)dλ = R(B; λ0 )Am+1 .
2πi ∂Ω
19
Hence
Am = (B − λ0 I)−1 Am+1 , m = 0, ±1, ±2, . . .
This recursion formula shows that (2.17) follows from the case m = 0. We
thus have to prove that
1 Z
− R(B; λ)dλ = I.
2πi ∂Ω
According to Proposition 2.27 it will suffice to prove that
1 Z
− R(B; λ)dλ = I
2πi |λ|=r
for sufficiently large r > 0. If |λ| = r > kBk we have the following absolutely
and uniformly convergent expansion
∞
(B − λI)−1 = − λ−n−1 B n
X
n=0
n 1 Z
λn R(B; λ)dλ, ∀n ∈ IN {0}.
[
B =−
2πi ∂Ω
PN m
Proof: It is sufficient to represent p(λ) in the form p(λ) = m=0 cm (λ − λ0 ) ,
λ0 6∈ Cl(Ω), and apply the last lemma. 2
20
where rf > r(B). Then Definitions 2.24 and 2.26 give the same result.
R(B; λ) R(B; µ)
R(B; λ)R(B; µ) = − − , λ 6= µ,
µ−λ λ−µ
we obtain
! !
1 Z 1 Z
f (B)g(B) = − f (λ)R(B; λ)dλ − g(µ)R(B; µ)dµ =
2πi ∂Ωf 2πi ∂Ωg
!
1 Z 1 Z
− f (λ) − g(µ)R(B; λ)R(B; µ)dµ dλ =
2πi ∂Ωf 2πi ∂Ωg
!
1 Z 1 Z g(µ)
− f (λ)R(B; λ) dµ dλ −
2πi ∂Ωf 2πi ∂Ωg µ − λ
!
1 Z 1 Z g(µ)
− f (λ) R(B; µ)dµ dλ.
2πi ∂Ωf 2πi ∂Ωg λ − µ
21
Since λ ∈ Ωg for any λ ∈ ∂Ωf and µ 6∈ Cl(Ωf ) for any µ ∈ ∂Ωg (see (2.19)),
the Cauchy theorem implies
1 Z g(µ) 1 Z f (λ)
dµ = g(λ), dλ = 0.
2πi ∂Ωg µ − λ 2πi ∂Ωf λ − µ
Therefore
1 Z
f (B)g(B) = − f (λ)g(λ)R(B; λ)dλ −
2πi ∂Ωf
!
1 Z 1 Z f (λ)
dλ g(µ)R(B; µ)dµ = (f g)(B) − 0 = (f g)(B). 2
2πi ∂Ωg 2πi ∂Ωf λ − µ
22
2.34. Theorem (the spectral mapping theorem) Let f be analytic
in some neighbourhood of σ(B). Then
23
2.35. Theorem Let f be analytic in some neighbourhood ∆f of σ(B) and g
be analytic in some neighbourhood ∆g of f (σ(B)). Then for the composition
g ◦ f (defined by (g ◦ f )(λ) := g(f (λ))) we have
Proof: The RHS of the last equality is well defined due to the spectral map-
ping theorem.
Let us take admissible sets Ωf and Ωg such that
σ(B) ⊂ Ωf ⊂ Cl(Ωf ) ⊂ ∆f ,
1 Z 1 Z
g(f (B)) = − g(ζ)R(f (B); ζ)dζ = − g(ζ)hζ (B)dζ =
2πi ∂Ωg 2πi ∂Ωg
!
1 Z 1 Z 1
− g(ζ) − R(B; λ)dλ dζ =
2πi ∂Ωg 2πi ∂Ωf f (λ) − ζ
!
1 Z 1 Z g(ζ)
− R(B; λ) dζ dλ =
2πi ∂Ωf 2πi ∂Ωg ζ − f (λ)
1 Z 1 Z
− g(f (λ))R(B; λ)dλ = − (g ◦ f )(λ)R(B; λ)dλ =
2πi ∂Ωf 2πi ∂Ωf
(g ◦ f )(B). 2
24
Summary
Ax = y, y ∈ Y. (2.22)
g(y) = g( lim Axn ) = lim g(Axn ) = lim (A∗ g)(xn ) = lim 0(xn ) = 0. 2
n→∞ n→∞ n→∞ n→∞
25
2.37. Corollary (a necessary condition of solvability) If (2.22) has a
solution then g(y) = 0, ∀g ∈ Ker(A∗ ).
Now we are going to prove that (2.23) is in fact an equality. We start with
the following corollary of the Hahn–Banach theorem.
Proof: Let
Y1 := {λy0 + y : λ ∈ IF, y ∈ Y0 }.
Define a functional g1 : Y1 → IF by the equality
g1 (λy0 + y) = λ.
2.39. Theorem
26
Suppose the contrary: there exists y0 ∈ N (A), y0 6∈ Ran(A). Lemma 2.38
implies the existence of g ∈ Y ∗ such that g(y0 ) = 1, g(y) = 0 for all y ∈
Ran(A). Then
(A∗ g)(x) = g(Ax) = 0, ∀x ∈ X,
i.e A∗ g = 0, i.e. g ∈ Ker(A∗ ). Since y0 ∈ N (A), we have g(y0 ) = 0. Contra-
diction! 2
Projections
2.41. Definition Let X be a normed space. An operator B ∈ B(X) is
called a projection if it is idempotent, i.e. B 2 = B.
Proof: Q2 = (I − P )2 = I − 2P + P 2 = I − 2P + P = I − P = Q, i.e. Q is a
projection.
P Q = P (I − P ) = P − P 2 = 0 = (I − P )P = QP.
27
2.44. Proposition Let L, L1 , L2 be subspaces of a vector space X. Then
L = L1 ⊕ L2 iff every element y of L can be uniquely written as y = y1 + y2
with yk ∈ Lk .
Proof: Exercise. 2
28
T S
and ∆ ∆0 = ∅. Let us consider the function f : ∆1 := ∆ S∆0 → C which equals S 1 in
∆ and 0 in ∆0 . It is analytic in ∆1 . Since σ(B) ⊂ Ω1 := Ω Ω0 and ∂Ω1 = ∂Ω ∂Ω0 ,
we have Z Z
1 1
P =− R(B; λ)dλ = − f (λ)R(B; λ)dλ = f (B).
2πi ∂Ω 2πi ∂Ω1
It is clear that f 2 = f . Hence P 2 = P (Theorem 2.31). It follows from the spectral
mapping theorem (Theorem 2.34) that σ(P ) = σ(f (B)) = {0, 1}. Thus P is a non–trivial
projection which commutes with g(B) for any function g analytic in some neighbourhood
of σ(B) (Corollary 2.32). P is called the spectral projection corresponding to the spectral
set σ. Since
g(B)P x = P g(B)x ∈ Ran(P ), ∀x ∈ X,
and
P g(B)x0 = g(B)P x0 = g(B)0 = 0, ∀x0 ∈ Ker(P ),
Ran(P ) and Ker(P ) are invariant under g(B), i.e.
Compact operators
2.47. Definition Let X be a normed space. A set K ⊂ X is relatively com-
pact if every sequence in K has a Cauchy subsequence. K is called compact
if every sequence in K has a subsequence which converges to an element of K.
Proof: See e.g. C. Goffman & G. Pedrick, First Course in Functional Anal-
ysis, Section 1.11, Lemma 1. 2
29
2.50. Proposition Let X be finite–dimensional. K ⊂ X is relatively
compact iff K is bounded. K ⊂ X is compact iff K is closed and bounded.
SX := {x ∈ X : kxk ≤ 1} (2.25)
Com(X, Y ) ⊂ B(X, Y ).
30
of a relatively compact set is relatively compact (see Proposition 2.48) we
conclude from T (W ) ⊂ T (tSX ) that T (W ) is relatively compact. 2
31
For a given ε > 0, we choose k so large that
ε
kT − Tk k < ,
3M
where M := supn∈IN kxn k. Then we determine n0 so that
ε
kTk x(n) (m)
n − Tk xm k < , ∀n, m ≥ n0 .
3
Now we have
kT x(n) (m)
n − T xm k < ε, ∀n, m ≥ n0 .
Hence (T x(n)
n ) is a Cauchy sequence, i.e. T ∈ Com(X, Y ). 2
Propositions (i) and (iii) of the last theorem mean that Com(X, Y ) is a
closed linear subspace of B(X, Y ). Let Y = X. It follows from (ii) that
Com(X) := Com(X, X)
is actually what we call an ideal in the Banach algebra B(X).
2.55. Definition We say that T ∈ B(X, Y ) is a finite rank operator if
Ran(T ) is finite–dimensional.
It is clear that any finite rank operator is compact (see Proposition 2.50).
Theorem 2.54(iii) implies that a limit of a convergent sequence of finite rank
operators is compact.
2.56. Theorem Every bounded subset of a normed space X is relatively
compact iff X is finite–dimensional.
Proof: See e.g. L.A. Liusternik & V.J. Sobolev, Elements of Functional
Analysis, §16). 2
2.57. Corollary The identity operator IX is compact iff X is finite–
dimensional. 2
2.58. Theorem Let T ∈ Com(X, Y ) and let at least one of the spaces
X and Y be infinite–dimensional. Then T is not invertible.
Proof: Suppose T is invertible: T −1 ∈ B(Y, X). Then operators T −1 T = IX
and T T −1 = IY are compact by Theorem 2.54(ii). This however contradicts
Corollary 2.57. 2
32
2.59. Corollary Let X be an infinite–dimensional Banach space and
T ∈ Com(X). Then 0 ∈ σ(T ). 2
2.60. Theorem Any non-zero eigenvalue λ of a compact operator T ∈
Com(X) has finite multiplicity, i.e. the subspace Xλ spanned by eigenvec-
tors of T corresponding to λ is finite–dimensional.
Proof: See e.g. W. Rudin, Functional Analysis, Theorem 4.25 and T. Kato,
Perturbation Theory For Linear Operators, Ch. III, §6, Section 5. 2
33
2.63. Proposition Let (Z, k · kZ ) be compactly embedded in (X, k · kX ),
Y be a normed space and let A : Y → (Z, k · kZ ) be bounded. Then
A ∈ Com(Y, X).
C m ([a, b]) is a Banach space. C m ([a, b]) is compactly embedded in C n ([a, b])
if m > n. (Let X and Z be some spaces of functions defined on a compact
set K. Normally it is reasonable to expect Z to be compactly embedded in
X if Z consists of functions which are “more smooth” than functions from
X.) It follows from Proposition 2.63 that if a linear operator T is continuous
from C n ([a, b]) to C m ([a, b]), m > n, then it is compact in C n ([a, b]). For
example the operator defined by the formula
Z b
(T u)(t) := k(t, s)u(s)ds, k ∈ C 1 ([a, b]2 ),
a
is bounded from C ([a, b]) = C([a, b]) to C 1 ([a, b]) and hence compact in
0
C([a, b]).
34
(3.1)–(3.3) imply
3.3. Theorem Every inner product space becomes a normed space by set-
ting kxk := (x, x)1/2 .
35
If IF = C then
So, every Hilbert Space is a Banach space and the whole Banach space theory
can be applied to Hilbert spaces.
Orthogonal complements
3.12. Theorem Let L be a closed linear subspace of a Hilbert space H.
For each x ∈ H there exists a unique point y ∈ L such that
(x − y, z) = 0, ∀z ∈ L.
36
Proof: Step I. Let yn ∈ L be such that kx−yn k → d := d(x, L). Applying the parallelogram
law to the vectors x − yn and x − ym and using the fact that (yn + ym )/2 ∈ L we have
Hence
(x − y, z) = λ = 0, ∀z ∈ L s.t. kzk = 1.
3.13. Remark The last theorem and its proof remain true in the case
when H is an inner product space and L is its complete linear subspace.
Proof: Exercise. 2
37
have H = M ⊕ M ⊥ .
If X is a normed space than the closed linear span of A is the closure of linA.
We will use the following notation for the closed linear span of A
spanA := Cl(linA).
38
3.20. Definition A system {eα }α∈J ⊂ H is called linearly independent if for
any finite subset {eα1 , . . . , eαN }, N
n=1 cn eαn = 0 =⇒ c1 = · · · = cN = 0.
P
for any m = 1, . . . , N . 2
Therefore kx − N 2 2
n=1 cn en k is minimal when cn = (x, en ) and kwk = kxk −
P
PN 2
n=1 |(x, en )| . The fact that w is orthogonal to L follows from Theorem
39
3.12 (see also Remark 3.13). Let us give an independent direct proof of this
fact:
N N
!
X X
(w, em ) = x − (x, en )en , em = (x, em ) − (x, en )(en , em ) =
n=1 n=1
N
(x, en )δnm = (x, em ) − (x, em ) = 0
X
(x, em ) −
n=1
Proof: Let ym = m
P
n=1 cn en . Then, using Pythagoras’
P∞
theorem (see Proposi-
2
tion 3.6) and the convergence of the series n=1 |cn | , we obtain
2
Xm
m
− yk k2 =
|cn |2 → 0, as k, m → ∞, (m > k).
X
kym cn en
=
n=k+1
n=k+1
n=1
40
for any N ∈ IN. Therefore
∞
|(x, en )|2 ≤ kxk2 . 2
X
n=1
(iii) (x, en ) = 0, ∀n ∈ IN =⇒ x = 0;
(iv) lin{en } is dense in H.
3.28. Examples
1. An orthonormal subset of a finite-dimensional Hilbert space is complete
iff it is a basis.
2. Let H = l2 ,
en = (0, . . . , 0, 1, 0, . . .).
| {z }
n
41
Then {en }n∈IN is a complete orthonormal set (why?).
3. The set {en }n∈Z ,
en (t) = ei2πnt ,
is orthonormal in the inner product space C([0, 1]) of Examples 3.2-3, 3.11-3
(check this!). It is one of the basic results of the classical Fourier analysis
that it has also the properties (ii)–(iv) of Theorem 3.26 (with Z and ∞
P
n=−∞
instead of IN and ∞
P P∞
n=1 correspondingly). The Fourier series n=−∞ (f, en )en
of a function f ∈ C([0, 1]) is convergent to f with respect to the norm k · k2 ,
but may be not uniformly convergent, i.e. not convergent with respect to the
norm k · k∞ .
lin{yn } = lin{en }.
Proof: Let us first construct a set {zn } such that lin{yn } = lin{zn } and zn
are linearly independent. We define zn inductively:
z1 = yn1 ,
z2 = yn2 ,
where yn2 is the first yn not in the linear span of z1 , and, generally,
zk = ynk ,
where ynk is the first yn not in the linear span of z1 , . . . , zk−1 . It is clear that
lin{yn } = lin{zn } and zn are linearly independent (prove this!).
Now we define inductively vectors en such that for any N the set {en }N n=1
is orthonormal and lin{en }N n=1 = lin{z } N
n n=1 . By our construction z1 is non-
zero, so
z1
e1 =
kz1 k
42
is well defined and has all the necessary properties in the case N = 1. Assume
that vectors e1 , . . . , eN −1 have been defined. By hypothesis zN ∈
/ LN −1 :=
N −1 N −1
lin{en }n=1 = lin{zn }n=1 , so the vector
N
X −1
uN = zN − (zN , ek )ek
k=1
Proof: Let H be a separable Hilbert space, {yn }n∈IN be a dense subset and
let {en } be the orthonormal set of the last theorem. Then lin{en } = lin{yn }
is dense in H, i.e. {en } is complete.
Conversely, assume that H contains a countable (or finite) complete or-
thonormal set {en }. Let
N
( )
X
QN = λn en : λ1 , . . . , λN ∈ Q + iQ ,
n=1
43
where Q is the field of rational numbers, and let
Q = ∪QN .
3.32. Theorem All infinite dimensional separable Hilbert spaces are iso-
morphic to l2 and hence to each other.
Proof: Let H be an infinite dimensional separable Hilbert space and let
{en }n∈IN be a complete orthonormal
P∞ set in it. Let us define U : l2 → H by
the formula U (λn )n∈IN = n=1 λn en (see Lemma 3.23). It is clear that U is
linear. Lemma 3.23 and Theorem 3.26 (Definition 3.27) imply that U is onto
and an isometry. 2
Any two complete orthonormal subsets of a Hilbert space H have the same cardinality.
This cardinality is called the dimension of H. Two Hilbert spaces are isomorphic iff they
have the same dimension. (For proofs see e.g. C. Goffman & G. Pedrick, First Course in
Functional Analysis, Section 4.7.).
44
Moreover, kf k = kzk.
Proof: It is clear that the equality (3.7) defines a bounded linear functional for an ar-
bitrary z ∈ H. So, we have to prove that any bounded linear functional on H has a unique
representation of the form (3.7).
Uniqueness. If z1 , z2 satisfy (3.7), then
(x, z1 ) = (x, z2 ), ∀x ∈ H,
i.e.
f (x)y − f (y)x ∈ Ker(f ), ∀x ∈ H.
Consequently
(f (x)y − f (y)x, y) = 0, ∀x ∈ H.
The left-hand side of the last equality equals f (x)kyk2 −f (y)(x, y) = f (x)kyk2 −(x, f (y)y).
Thus
f (y)
f (x) = (x, z), ∀x ∈ H, where z := y.
kyk2
The equality kf k = kzk follows from the relations
H1 3 x 7→ f (x) := (Bx, y) ∈ IF
45
is a linear functional and
H2 3 y 7→ B ∗ y := z ∈ H1
kB ∗ yk ≤ kBkkyk, ∀y ∈ H2 ,
Proof: (i)
((αB1 + βB2 )x, y) = α(B1 x, y) + β(B2 x, y) = α(x, B1∗ y) + β(x, B2∗ y) =
(x, (αB1∗ + βB2∗ )y), ∀x ∈ H1 , ∀y ∈ H2 .
(ii)
(T Bx, z) = (Bx, T ∗ z) = (x, B ∗ T ∗ z), ∀x ∈ H1 , ∀z ∈ H3 .
(iii)
46
Consequently
((B − B ∗∗ )x, y) = 0, ∀x ∈ H1 , ∀y ∈ H2 .
Taking y = (B − B ∗∗ )x we obtain (B − B ∗∗ )x = 0, ∀x ∈ H1 .
(iv) follows from (4.2) and (iii):
(v)
kB ∗ Bk ≤ kB ∗ kkBk = kBk2
(see (iv)). On the other hand
kBk2 = sup kBxk2 = sup (Bx, Bx) = sup (x, B ∗ Bx) ≤
kxk=1 kxk=1 kxk=1
BB −1 = IH2 , B −1 B = IH1 ,
(see (ii)). 2
Parts (i)–(iii) of Theorem 4.3 show that B(H) is a Banach algebra with
an involution. Part (v) shows that it is, in fact, a C ∗ –algebra.
kx∗ xk = kxk2 , ∀x ∈ E,
is called a C ∗ –algebra.
47
Let us explain why this equality is different from that of Theorem 4.3(i). If
H1 and H2 are Hilbert spaces, then we have two definitions of the adjoint
of B ∈ B(H1 , H2 ): Definition 4.2 and that for operators acting on normed
spaces. B ∗ ∈ B(H2 , H1 ) according to the first one, while the second one gives
us an operator B ∗ ∈ B(H2∗ , H1∗ ). In order to connect them to each other we
have to use the identification H1∗ ∼ H1 , H2∗ ∼ H2 given by Theorem 3.34.
This identification is conjugate linear (anti-linear):
This is why we have the complex conjugation in Theorem 4.3(i) and do not
have it in (4.3).
Note also that we use the following definition of linear operations on X ∗ (for
an arbitrary normed space X)
In this case Theorem 4.3(i) is valid for operators acting on normed spaces.
Proof:
B ∗ y = 0 ⇐⇒ (x, B ∗ y) = 0, ∀x ∈ H1 ⇐⇒ (Bx, y) = 0, ∀x ∈ H1 ⇐⇒
y ∈ Ran(B)⊥ .
48
Proof: Note that for any linear subspace M of a Hilbert space H we have
the equality Cl(M ) = M ⊥⊥ . (Prove this!) 2
Proof: We have
kBzk2 = (Bz, Bz) = (B ∗ Bz, z),
kB ∗ zk2 = (B ∗ z, B ∗ z) = (BB ∗ z, z)
for any z ∈ H. So, (4.4) holds if B is normal.
If kBzk = kB ∗ zk, ∀z ∈ H, then using the polarization identity (see Propo-
sition 3.7) we obtain
i.e.
((B ∗ B − BB ∗ )x, y) = 0, ∀x, y ∈ H.
Taking y = (B ∗ B − BB ∗ )x we deduce that (B ∗ B − BB ∗ )x = 0, ∀x ∈ H. 2
49
to each other.
Proof: (i) follows from Theorems 4.4 and 4.7. Applying (i) to B − αI in
place of B we obtain (ii). (Exercise: prove that B − αI is normal.)
Finally, suppose Bx = αx, By = βy and α 6= β. We have from (ii)
Consequently U ∗ U = IH1 (why?). On the other hand (iii) implies that U is a linear isom-
etry of H1 onto H2 . Hence U is invertible. Since U ∗ U = IH1 , we have U −1 = U ∗ (why?),
i.e. U is unitary. 2
Theorem 4.9 shows that the notion of a unitary operator coincides with the
notion of an isomorphism of Hilbert spaces (see Definition 1.4) and that these
isomorphisms preserve the inner products (see also the equality in the para-
graph between Theorems 3.31 and 3.32).
4.10. Theorem
(i) B is self-adjoint, λ ∈ IR =⇒ λB is self-adjoint.
(ii) B1 , B2 are self-adjoint =⇒ B1 + B2 is self-adjoint.
(iii) Let B1 , B2 be self-adjoint. Then B1 B2 is self-adjoint if and only if B1
and B2 commute.
(iv) If Bn , n ∈ IN are self-adjoint and kB − Bn k → 0, then B is self-adjoint.
50
Proof: Exercise. 2
(check this expanding the right-hand side; this is the polarization identity for
operators) and similarly
Since
(Bz, z) = (Bz, z) = (z, Bz), ∀z ∈ H,
we have
(Bx, y) = (x, By), ∀x, y ∈ H,
i.e. B is self-adjoint. 2
(Bx, x)
λ= ∈ IR. 2
kxk2
51
4.13. Theorem Let H be a Hilbert space. Each of the following four
properties of a projection P ∈ B(H) implies the other three:
(i) P is self-adjoint.
(ii) P is normal.
(iii) Ran(P ) = Ker(P )⊥ .
(iv) (P x, x) = kP xk2 , ∀x ∈ H.
Proof: It is trivial that (i) implies (ii).
Theorem 4.8(i) shows that Ker(P ) = Ran(P )⊥ if P is normal. Since P is a projection
Ran(P ) is closed (see Lemma 2.45). So, Ran(P ) = Ran(P )⊥⊥ = Ker(P )⊥ (see the proof
of Corollary 4.5). Thus (ii) implies (iii).
Let us prove that (iii) implies (i). Indeed, if (iii) holds, then
i.e. P is self-adjoint. So, we have proved that the properties (i)–(iii) are equivalent to each
other.
Now it is sufficient to prove that (iii) is equivalent to (iv). We have seen above that (iii)
implies the equality (P x, (I − P )x) = 0, ∀x ∈ H. Hence
(P x, x) = (P x, P x) + (P x, (I − P )x) = (P x, P x) = kP xk2 , ∀x ∈ H.
Finally, assume (iv) holds. Let us take arbitrary y ∈ Ran(P ) and z ∈ Ker(P ) and consider
x = y + tz, t ∈ IR. It is clear that P x = y. Consequently
Thus
t(y, z) ≥ −kyk2 , ∀t ∈ IR,
i.e.
(y, z) = 0, ∀y ∈ Ran(P ), ∀z ∈ Ker(P ),
52
by saying that P is an orthogonal projection.
Numerical range
4.15. Theorem Let H be a Hilbert space and B ∈ B(H). Suppose there
exists c > 0 such that
ckxk2 ≤ kBxkkxk, ∀x ∈ H,
i.e.
kxk ≤ c−1 kBxk, ∀x ∈ H. (4.7)
Let us prove that Ran(B) is closed. For any y ∈ Cl(Ran(B)) there exist
xn ∈ H such that Bxn → y as n → ∞. It follows from (4.7) that
Hence (xn ) is a Cauchy sequence in the Hilbert space H. Let us denote its
limit by z. We have
Bz = B lim xn = lim Bxn = y.
n→∞ n→∞
53
is called the numerical range of the operator B.
It is clear that
so,
Num(B) ⊂ {µ ∈ C : |µ| ≤ kBk}. (4.9)
The numerical range of any bounded linear operator is convex (see e.g. G.
Bachman & L. Narici, Functional Analysis, 21.4).
54
Spectra of self-adjoint and unitary operators
It follows from Theorem 4.11 that if B ∈ B(H) is self-adjoint then Num(B) ⊂
IR. Therefore Theorem 4.17 implies the following result.
U − λI = U (I − λU −1 )
and kλU −1 k = |λ| < 1. Hence U − λI is invertible by Lemmas 2.3(i) and 2.4.
So, λ ∈
/ σ(U ), i.e. σ(U ) ⊂ {λ ∈ C : |λ| = 1}. 2
kB 2 xk = kB ∗ Bxk, ∀x ∈ H.
55
4.21. Theorem Let B ∈ B(H) be a normal operator. Then
Pn y = z, where y = z + w, z ∈ Ln , w ∈ L⊥
n,
It is clear that any Banach space having a Schauder basis is separable (cf.
the proof of Theorem 3.31). Theorem 4.22 remains valid if we replace H by
56
a Banach space having a Schauder basis. In 1973 P. Enflo gave a negative
solution to the long standing problem on existing of a Schauder basis in any
separable Banach space: he constructed a separable Banach space without
Schauder basis. He also proved that there exist a separable Banach space
and a compact operator acting in it such that this operator is not a limit of
a convergent sequence of finite rank operators.
Hilbert–Schmidt operators
Let H1 and H2 be separable Hilbert spaces and {en }n∈IN be a complete or-
thonormal set in H1 . Then for any operator T ∈ B(H1 , H2 ) the quantity
P∞ 2
n=1 kT en k is independent of the choice of a complete orthonormal set {en }
(this quantity may be infinite). Indeed, if {fn }n∈IN is a complete orthonormal
set in H2 then
∞ ∞ X
∞ ∞ X
∞ ∞
∗
2 2 2
kT ∗ fk k2 .
X X X X
kT en k = |(T en , fk )| = |(en , T fk )| =
n=1 n=1 k=1 n=1 k=1 k=1
So, ∞
P∞
2
n=1 kT en k = kT e0n k2 for any two complete orthonormal sets {en }
P
n=1
and {e0n } in H1 .
and the operators for which it is finite are called Hilbert–Schmidt operators.
Note that
kT k ≤ kT kHS . (4.12)
Indeed,
∞
!
∞
∞
X
X
X
kT xk =
T (x, en )en
=
(x, en )T en
≤ |(x, en )|kT en k ≤
n=1 n=1 n=1
∞
!1/2 ∞ !1/2
|(x, en )|2 kT en k2
X X
= kT kHS kxk, ∀x ∈ H1 ,
n=1 n=1
57
(see Theorem 3.26).
and therefore
1/2
∞
kT en k2
X
kT − TN k ≤ → 0 as N → ∞.
n=N +1
4.27. Example (Integral operators) Let H = L2 ([0, 1]) and let a mea-
surable function
k : [0, 1] × [0, 1] → IF
be such that Z 1 Z 1
|k(t, τ )|2 dτ dt < ∞.
0 0
Define the operator K by
Z 1
(Kf )(t) := k(t, τ )f (τ )dτ.
0
58
Then K : L2 ([0, 1]) → L2 ([0, 1]) is a Hilbert–Schmidt operator and
Z 1 Z 1 1/2
2
kKkHS = |k(t, τ )| dτ dt .
0 0
Proof: For t ∈ [0, 1] let kt (τ ) := k(t, τ ). Let {en }n∈IN be a complete orthonor-
mal set in L2 ([0, 1]). Then
Z 1
Ken (t) = k(t, τ )en (τ )dτ = (kt , en ).
0
Hence Z 1 Z 1
2 2
kKen k = |Ken (t)| dt = |(kt , en )|2 dt
0 0
and therefore
∞ ∞ Z 1 Z ∞
1 X
kKen k2 = |(kt , en )|2 dt = |(kt , en )|2 dt =
X X
Proof: This result follows from Theorems 2.61 and 4.20. Since we have
not proved Theorem 2.61, we give an independent proof of Theorem 5.1.
There is nothing to prove if T = 0. So, let us suppose that T 6= 0.
It follows from Theorem 4.21 that there exist xn ∈ H, n ∈ IN such that
kxn k = 1 and |(T xn , xn )| → kT k. We can suppose that the sequence of
numbers (T xn , xn ) is convergent (otherwise we could take a subsequence of
59
(xn )). Let λ be the limit of this sequence. It is clear that |λ| = kT k =
6 0. We
have
Hence kT xn − λxn k → 0 as n → ∞.
Since T is a compact operator the sequence (T xn ) has a Cauchy subsequence
(T xnk )k∈IN . The last sequence is convergent because the space H is complete.
Let us denote its limit by y. It follows from the formula
1
xnk = (T xnk − (T xnk − λxnk ))
λ
that the sequence (xnk )k∈IN converges to x := λ−1 y. Consequently
Tx = T lim xnk = lim T xnk = y = λx
k→∞ k→∞
and
kxk =
lim xnk
= lim kxnk k = 1.
k→∞ k→∞
60
|λ1 | ≥ |λ2 | ≥ · · · (5.4)
and
lim λn = 0, if N = ∞. (5.5)
n→∞
Lk := lin{e1 , . . . , ek }, Hk := L⊥
k.
It is clear that
L1 ⊂ L2 ⊂ · · · ⊂ Lk , H1 ⊃ H2 ⊃ · · · ⊃ Hk . (5.6)
Pk
Any x ∈ Lk has a unique representation of the form x = n=1 cn en , cn ∈ C.
Consequently
k k
λn cn en ∈ Lk , T ∗ x =
X X
Tx = λn cn en ∈ Lk
n=1 n=1
T Lk ⊂ Lk , T ∗ Lk ⊂ Lk .
For any z ∈ Hk = L⊥
k we have
(x, T z) = (T ∗ x, z) = 0, (x, T ∗ z) = (T x, z) = 0, ∀x ∈ Lk ,
T Hk ⊂ Hk , T ∗ Hk ⊂ Hk .
Tk := T |Hk ∈ B(Hk ).
It is easy to see that Tk is a compact normal operator (why?). There are two
possibilities.
61
Case I. Tk = 0. In this case the construction terminates.
Case II. Tk 6= 0. In this case there exists an eigenvalue λk+1 6= 0 of Tk
such that |λk+1 | = kTk k = kT |Hk k (see Theorem 5.1). It follows from the
construction and from (5.6) that |λk+1 | ≤ |λk | ≤ · · · ≤ |λ1 |. Let ek+1 ∈ Hk =
L⊥k be an eigenvector of Tk corresponding to λk+1 and such that kek+1 k =
1. It is clear that ek+1 is an eigenvector of T and that the set {en }k+1 n=1 is
orthonormal.
This construction gives us a finite or a countable orthonormal set {en }N n=1 ,
N ∈ IN ∪ {∞}, of eigenvectors of T and corresponding non-zero eigenvalues
λn satisfying (5.4). This set is finite if we have Case I for some k and infinite
if we always have Case II.
Step II Let us prove (5.5). Suppose (5.5) does not hold. Then there exists
δ > 0 such that |λn | ≥ δ, ∀n ∈ IN (see (5.4)). Consequently
HN := ∩N
n=1 Hn .
HN = Ker(T ). (5.7)
62
(cf. the proof of Theorem 3.22). Thus x has a representation of the form
(5.1), where cn = (x, en ) and
N
X
y := x − (x, en )en ∈ HN = Ker(T ).
n=1
Taking the inner products of both sides of (5.1) with ek and using (5.7) we
obtain ck = (x, ek ), i.e. a representation of the form (5.1) is unique.
Step V (5.2) follows immediately from (5.1) and the continuity of T . It
N
is left to prove (5.3). Let us take an arbitrary λ ∈ C\ {λn }n=1 ∪ {0} . It is
clear that the distance d from λ to the closed set {λn }N n=1 ∪ {0} is positive.
(5.1), (5.2) imply
N
X
(T − λI)x = (λn − λ)(x, en )en − λy.
n=1
Note that
N
k(T − λI)−1 xk2 = |λn − λ|−2 |(x, en )|2 + |λ|−2 kyk2 ≤
X
n=1
N
!
−2 2 2
= d−2 kxk2 < ∞
X
d |(x, en )| + kyk
n=1
(see Proposition 3.6, Lemma 3.23 and (5.7)). So, we have proved that λ ∈ /
σ(T ), i.e. σ(T ) ⊂ {λn }N
n=1 ∪{0}. On the other hand we have {λ } N
n n=1 ⊂ σ(T ).
2
63
Let Pn be the orthogonal projection onto the one–dimensional subspace gen-
erated by en , i.e. Pn := (·, en )en and let PKer (T) be the orthogonal projection
onto Ker(T ). Then (5.1), (5.2) and (5.8) can be rewritten in the following
way:
N
X
I= Pn + PKer (T) , (5.9)
n=1
N
X
T = λ n Pn , (5.10)
n=1
N
R(T ; λ) = (T − λI)−1 = (λn − λ)−1 Pn − λ−1 PKer (T)
X
(5.11)
n=1
(see Step IV of the proof of Theorem 5.2), where the series are strongly
convergent. (A series ∞ n=1 An , An ∈ B(X, Y ), is called strongly convergent
P
P∞
if the series n=1 An x converges in Y for any x ∈ X.) On the other hand it
is easy to see that in the case N = ∞ the series (5.9) and (5.11) do not
converge in the B(H)–norm, while (5.10) does (prove this!).
Suppose a function f is analytic in some neighbourhood ∆f of σ(T ) and
Ω is an admissible set such that
σ(T ) ⊂ Ω ⊂ Cl(Ω) ⊂ ∆f .
Then we obtain from Definition 2.26, (5.11) and the Cauchy theorem
1 Z
f (T )x = − f (λ)R(T ; λ)dλ x =
2πi ∂Ω
N
!
1 Z
(λn − λ)−1 Pn x − λ−1 PKer (T) x dλ =
X
− f (λ)
2πi ∂Ω n=1
N
1 Z
f (λ)(λn − λ)−1 dλ Pn x +
X
−
n=1 2πi ∂Ω
1 Z
f (λ)λ−1 dλ PKer (T) x =
2πi ∂Ω
N
X
f (λn )Pn x + f (0)PKer (T) x, ∀x ∈ H. (5.12)
n=1
64
Note that if dim(H) < +∞ it may happen that 0 6∈ σ(T ) and 0 6∈ Ω. If in
this case f (0) 6= 0 then
1 Z
f (λ)λ−1 dλ = 0 6= f (0).
2πi ∂Ω
Nevertheless (5.12) holds because PKer (T) = 0, since Ker (T) = 0.
It is clear that the RHS of (5.12) is well defined for any bounded on σ(T )
not necessarily analytic function f . This motivates the following definition.
5.4. Definition Let T ∈ Com(H) be a normal operator and f be a bounded
function on σ(T ). Then
N
X
f (T )x := f (λn )Pn x + f (0)PKer (T) x, ∀x ∈ H.
n=1
Note that if dim(H) < +∞ and 0 6∈ σ(T ), then f may be not defined
at 0. The above definition, however, still makes sense because in this case
PKer (T) = 0, since Ker (T) = 0, and we assume that f (0)PKer (T) = 0.
The operator f (T ) is well defined because
2
N
X
f (λn )Pn x + f (0)PKer (T) x
=
n=1
N
|f (λn )(x, en )|2 + |f (0)|2 kPKer (T) xk2 ≤
X
n=1
!2
sup |f (λ)| kxk2 < ∞, ∀x ∈ H, (5.13)
λ∈σ(T )
(cf. Step V of the proof of Theorem 5.2). It follows from (5.13) that
kf (T )k ≤ sup |f (λ)|.
λ∈σ(T )
65
Since kf (T )en k = kf (λk )en k = |f (λn )| and kf (T )yk = kf (0)yk = |f (0)|kyk
for any y ∈ Ker(T ), we have
(
supλ∈σ(T ) |f (λ)| if Ker (T) 6= {0},
kf (T )k ≥
supλ∈σ(T )\{0} |f (λ)| if Ker (T) = {0}.
Thus (
supλ∈σ(T ) |f (λ)| if Ker (T) 6= {0},
kf (T )k = (5.14)
supλ∈σ(T )\{0} |f (λ)| if Ker (T) = {0}.
If dim(H) < +∞ and Ker(T ) = {0}, then 0 6∈ σ(T ) and σ(T ) \ {0} =
σ(T ). If dim(H) = +∞ and Ker(T ) = {0}, then 0 ∈ σ(T ) is the only limit
point of the set σ(T ) \ {0}. Therefore
sup |f (λ)| = sup |f (λ)|, if f is continuous at 0.
λ∈σ(T ) λ∈σ(T )\{0}
Proof: Exercise. 2
66