Sie sind auf Seite 1von 77

Contents

1 Banach spaces 7
1.1 Metric spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2 Normed spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.3 Hölder and Minkowski inequality . . . . . . . . . . . . . . . . . . 22

2 Bounded maps; the dual space 25


2.1 Bounded linear maps . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.2 The dual space and the Hahn-Banach theorem . . . . . . . . . . 31
2.3 Examples of dual spaces . . . . . . . . . . . . . . . . . . . . . . . 36
2.4 The Banach space adjoint and the bidual . . . . . . . . . . . . . 38

3 Linear operators in Banach spaces 45


3.1 Baire’s theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.2 Uniform boundedness principle . . . . . . . . . . . . . . . . . . . 46
3.3 The open mapping theorem . . . . . . . . . . . . . . . . . . . . . 55
3.4 The closed graph theorem . . . . . . . . . . . . . . . . . . . . . . 57
3.5 Projections in Banach spaces . . . . . . . . . . . . . . . . . . . . 61
3.6 Weak convergence . . . . . . . . . . . . . . . . . . . . . . . . . . 62

4 Hilbert spaces 67
4.1 Hilbert spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
4.2 Orthogonality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.3 Orthonormal systems . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.4 Linear operators in Hilbert spaces . . . . . . . . . . . . . . . . . 80
4.5 Projections in Hilbert spaces . . . . . . . . . . . . . . . . . . . . 84
4.6 The adjoint of an unbounded operator . . . . . . . . . . . . . . . 87

5 Spectrum of linear operators 97


5.1 The spectrum of a linear operator . . . . . . . . . . . . . . . . . 97
5.2 The resolvent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
5.3 The spectrum of the adjoint operator . . . . . . . . . . . . . . . . 106
5.4 Compact operators . . . . . . . . . . . . . . . . . . . . . . . . . . 109
5.5 Hilbert-Schmidt operators . . . . . . . . . . . . . . . . . . . . . . 121
5.6 Polar decomposition . . . . . . . . . . . . . . . . . . . . . . . . . 124

3
4 CONTENTS CONTENTS 5

A Lp spaces 129 Notation


A.1 A reminder on measure theory . . . . . . . . . . . . . . . . . . . 130
A.2 Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133 The letter K usually denotes either the real field R or the complex field C. The
A.3 Lp spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136 positive real numbers are denoted by R+ := (0, ∞).

B Exercises 139

References 152
Chapter 1. Banach spaces 7

Chapter 1

Banach spaces

1.1 Metric spaces


We repeat the definition of a metric space.

Definition 1.1. A metric space (M, d) is a non-empty set M together with a


map

d:M ×M →R

such that for all x, y, z ∈ M :

(i) d(x, y) = 0 ⇐⇒ x = y,
(ii) d(x, y) = d(y, x),
(iii) d(x, y) ≤ d(x, z) + d(z, y).

The last inequality is called triangle inequality. Usually the metric space (M, d)
is denoted simply by M .

Note that the triangle inequality together with the symmetry of d implies

d(x, y) ≥ 0, x, y ∈ M,

since 0 = d(x, x) ≤ d(x, y) + d(y, x) = 2d(x, y).


It is easy to check that

|d(x, y) − d(y, z)| ≤ d(x, z), x, y, z ∈ M.

A subset N ⊆ M is called bounded if

diam N := sup{d(x, y) : x, y ∈ N } < ∞.


8 1.1. Metric spaces Chapter 1. Banach spaces 9

Let r > 0 and x ∈ M . Then Remarks. • Every convergent sequence is a Cauchy sequence.
Br (x) := {y ∈ M : d(x, y) < r} =: open ball with centre x and radius r, • Every Cauchy sequence is bounded. Recall that a sequence (xn )n∈N is
bounded if the set {xn : n ∈ N} is bounded.
Kr (x) := {y ∈ M : d(x, y) ≤ r} =: closed ball with centre x and radius r,
Sr (x) := {y ∈ M : d(x, y) = r} =: sphere with centre x and radius r. Not every metric space is complete, but every metric space can be completed in
the following sense.
Examples. • R with the d(x, y) = |x − y| is a metric space.
• Let X be a set and define d : X × X → R by d(x, y) = 0 for x = y Definition 1.6. Let (M, dM ) and (N, dN ) be metric spaces. A map f : M → N
and d(x, y) = 1 for x 6= y. Then (X, d) is a metric space. d is called the is called an isometry if and only if dN (f (x), f (y)) = dM (x, y) for all x, y ∈ M .
discrete metric on X. The spaces M and N are called isometric if there exists a bijective isometry
f : M → N.
Let (M, d) be a metric space. Recall that the metric d induces a topology on
M : a set U ⊆ M is open if and only if for every p ∈ U there exists an ε > 0 Note that an isometry is necessarily injective since x 6= y implies f (x) 6= f (y)
such that Bε (p) ⊆ U . In particular, the open balls are open and closed balls are because d(f (x), f (y)) = d(x, y) 6= 0, and that every isometry is continuous.
closed subsets of M . Let x ∈ M . A subset U ⊆ M is called a neighbourhood of
x if there exists an open set Ux such that x ∈ Ux ⊆ U . Theorem 1.7. Let (M, d) be a metric space. Then there exists a complete
It is easy to see that the topology generated by d has the Hausdorff property, metric space (M b and an isometry ϕ : M → M
c, d) c such that ϕ(M ) = M
c. M
c is
that is, for every x 6= y ∈ M there exist neighbourhoods Ux of x and Uy of y called completion of M ; it is unique up to isometry.
with Ux ∩ Uy = ∅.
Recall that a set N ⊆ M is called dense in M if N = M , where N denotes the Proof. Let
closure of N .
CM := {(xn )n∈N ⊆ M : (xn )n∈N is a Cauchy sequence in M }
Definition 1.2. A sequence (xn )n∈N ⊆ M converges to x ∈ M if and only if
lim d(xn , x) = 0, that is, be the set of all Cauchy sequences in M . We define the equivalence relation ∼
n→∞ on CM by
∀ε > 0 ∃N ∈ N : n ≥ N =⇒ d(xn , x) < ε.
x∼y ⇐⇒ d(xn , yn ) → 0, n → ∞
The limit x is unique. A sequence (xn )n∈N is a Cauchy sequence in M if and
only if for all x = (xn )n∈N , y = (yn )n∈N ∈ CM . It is easy to check that ∼ is indeed a
equivalence relation (reflexivity and symmetry follow directly from properties (i)
∀ε > 0 ∃N ∈ N : m, n ≥ N =⇒ d(xn , xm ) < ε. and (ii) of the definition of a metric and transitivity of ∼ is a consequence of
the triangle inequality).
Definition 1.3. A metric space in which every Cauchy sequence is convergent, c := CM / ∼ the set of all equivalence classes. The equivalence class
is called a complete metric space. Let M
containing x = (xn )n∈N is denoted by [x]. On M c we define
Definition 1.4. Let (X, OX ) and (Y, OY ) be topological spaces.
db : M
c×M
c → R, b
d([x], [y]) = lim d(xn , yn ). (1.1)
(i) A function f : X → Y is called continuous if and only if f −1 (U ) is open n→∞

in X for every U open in Y .


We have to show that db is well-defined.
(ii) An bijective function f : X → Y is called a homeomorphism if and only Let (xn )n∈N ∈ [x] and (yn )n∈N ∈ [y]. Then
if f and f −1 are contiunous.
|d(xn , yn ) − d(xm , ym )| ≤ |d(xn , yn ) − d(xm , yn )| + |d(xm , yn ) − d(xm , ym )|
The following lemma is often useful. ≤ d(xn , xm ) + d(yn , ym ) → 0, m, n → ∞.

Lemma 1.5. Let (M, d) be a complete metric space and N ⊆ M . Then N is Since (d(xn , yn ))n∈N is a Cauchy sequence in the complete space R, the limit in
closed in M if and only if (N, d|M ) is complete. (1.1) exists.
10 1.1. Metric spaces Chapter 1. Banach spaces 11

Moreover, for (x̃n )n∈N ∈ [x] and (ỹn )n∈N ∈ [y] it follows that We have shown that ϕ(M ) is a dense subset of the complete metric space (M b
c, d)
and that ϕ is an isometry.
|d(xn , yn ) − d(x̃n , ỹn )| ≤ |d(xn , yn ) − d(x̃n , yn )| + |d(x̃n , yn ) − d(x̃n , ỹn )| Finally, we have to show that Mc is unique (up to isometry). Let (N, dN ) be
≤ d(xn , x̃n ) + d(yn , ỹn ) → 0, n → ∞. complete metric space and ψ : M → N an isometry such that ψ(M ) = N . Then
the map
Hence db is well-defined.
T : ϕ(M ) → ψ(M ), T (ϕ(x)) = ψ(x)
Let
c → N by
can be extended to a surjective isometry T : ϕ(M ) = M
c,
ϕ:M →M ϕ(x) = [(x)n∈N ].
T x = T ( lim xn ) := lim T xn
n→∞ n→∞
We will show that (M c, d)
b is a complete metric space, that ϕ is an isometry and
that ϕ(M ) = M c in several steps. for x = lim xn with xn ∈ ϕ(M ), n ∈ N.
n→∞
b is a metric space.
c, d)
Step 1: (M
c. Then Examples. • Cn with d(x, y) = max{|xj − yj | : j = 1, . . . , n} is a complete
Proof. Let [x], [y], [z] ∈ M metric space.
b p
• 0 = d([x], [y]) = lim d(xn , yn ) ⇐⇒ x∼y ⇐⇒ [x] = [y]. • Cn with d(x, y) = |x1 − y1 |2 + · · · + |xn − yn |2 is a complete metric
n→∞
b b space.
• d([x], [y]) = lim d(xn , yn ) = lim d(yn , xn ) = d([y], [x]).
n→∞ n→∞
b b • Let C([a, b]) be the set of all continuous functions on the interval [a, b].
• d([x], [y]) = lim d(xn , yn ) ≤ lim d(xn , zn ) + d(zn , yn ) = d([x], [z]) +
n→∞ n→∞ For f, g ∈ C([a, b]) let
b
d([z], [y]).
d1 (f, g) := max{|f (x) − g(x)| : x ∈ [a, b]},
Z b
Step 2: ϕ is an isometry.
Proof. This follows immediately from the definition. d2 (f, g) := |f (x) − g(x)| dx.
a
Step 3: ϕ(M ) = M c.
Then d1 and d2 are metrics on C([a, b]). (C([a, b]), d1 ) is complete, (C([a, b]), d2 )
Proof. Let (xn )n∈N ∈ [x] ∈ M c and ε > 0. Then there exists an N ∈ N such that
is not complete.
d(xn , xm ) < 2ε , m, n ≥ N . Let z := xN ∈ M . Then
ε Remark. The completion of (C([a, b]), d2 ) is L1 (a, b) (the set of all Lebesgue
b
d(ϕ(z), [x]) = lim d(xN , xn ) ≤ < ε.
n→∞ 2 integrable functions on (a, b)).

Next we show that (M b is complete. Let (x̂n )n∈N be a Cauchy sequence in


c, d) Definition 1.8. A metric space is called separable if it contains a countable
c. Since ϕ(M ) is dense in M
M c there exists a sequence z = (zn )n∈N ⊆ M such dense subset.
that
Proposition 1.9. Let (M, d) be a separable metric space and N ⊆ M . Then
b xn , zn ) < 1 ,
d(b n ∈ N. N is separable.
n
The sequence z is a Cauchy sequence in M because Proof. We have to show that there exists a countable set B ⊆ N such that B ⊇
N where the closure is taken with respect to the metric on M . By assumption
b
d(zn , zm ) = d(ϕ(z b b b
n ), ϕ(zm )) ≤ d(ϕ(zn ), x̂n ) + d(x̂n , x̂m ) + d(x̂m , ϕ(zm )) on M there exists a countable set A := {xn : n ∈ N} ⊆ M such that A = M .
1 1 1
b n , x̂m ) + Let J := {(n, m) ∈ N × N : ∃y ∈ N with d(xn , y) < m }. For every (n, m) ∈ J
< + d(x̂ → 0, m, n → ∞.
n m choose a yn,m ∈ N and define B := {yn,m : (n, m) ∈ J}. Obviously, B is a
countable subset of N . To show that B is dense in N it suffices to show that for
The sequence (x̂n )n∈N converges to [z] because
every y ∈ N and k ∈ N there exists a b ∈ B such that d(b, y) < k1 . By definition
1
1 of A there exists a xn ∈ A such that d(xn , y) < 2k . In particular, (n, 2k) ∈ J.
b n , z) ≤ d(x̂
d(x̂ b n , ϕ(zn )) + d(ϕ(z
b n ), z) < + lim d(zn , zm ) → 0, n → ∞.
n m→∞ It follows that d(yn,2k , y) ≤ d(yn,2k , xn ) + d(xn , y) < k1 .
12 1.2. Normed spaces Chapter 1. Banach spaces 13

1.2 Normed spaces For x ∈ X we denote the equivalence class of X/M containing x by [x]. Then
X/M is a vector space if we set
Definition 1.10. Let X be a vector space over K. A norm on X is a map
[x] + [y] := [x + y], α[x] := [αx], x, y ∈ X, α ∈ K.
k·k :X →R For x ∈ X let dist(x, M ) := inf{kx − mk : m ∈ M }.
such that for all x, y ∈ X, α ∈ K • (X/M, k · k∼ ) is a normed space with

(i) kxk = 0 ⇐⇒ x = 0, k · k∼ : X/M → R, k[x]k∼ := dist(x, M ).


(ii) kαxk = |α| kxk, Proof. First we show that k · k∼ is well-defined. For x, y ∈ X with x − y ∈
(iii) kx + yk ≤ kxk + kyk. M we find
∈M
z }| {
Remarks. • Note that the implication ⇐ in (i) follows from (ii) because dist(x, M ) = inf{kx − mk : m ∈ M } = inf{ky − (y − x +m)k : m ∈ M }
k0k = k2 · 0k = 2k0k.
= inf{ky − mk : m ∈ M } = dist(y, M ).
• Note that kxk ≥ 0 for all x ∈ X because 0 = kx − xk ≤ 2kxk. The last Property (ii) in the definition of a norm is easily checked. For property
inequality follows from the triangle inequality (iii) and (ii) with α = −1. (iii) let [x], [y] ∈ X/M . Then

Remark. A function [ · ] : X → R which satisfies only (ii) and (iii) of Defi- k[x] + [y]k∼ = k[x + y]k∼ = inf{kx + y − mk : m ∈ M }
nition 1.10 is called a seminorm. As seen in the remark above for norms, a = inf{kx − mx + y − my k : mx , my ∈ M }
seminorm is non-negative and satisfies [0] = 0. ≤ inf{kx − mx k : mx ∈ M } + inf{ky − my k : my ∈ M }
= k[x]k∼ + k[y]k∼ .
Remark. A norm on X induces a metric on X by setting
It is clear that [x] = 0 implies k[x]k∼ = 0. Now assume that k[x]k∼ = 0.
d(x, y) := kx − yk, x, y ∈ X. We have to show that x ∈ M . By definition of dist there exists a sequence
(mn )n∈N such that kx − mn k → 0, that is, (mn )n∈N converges to x. Since
Hence a norm induces a topology on X via the metric and we have the concept M is closed, it follows that x ∈ M .
of convergence etc. on a normed space.
• Let X be a Banach space and M a closed subspace. Then X/M is Banach
Definition 1.11. A complete normed space is called a Banach space. space with the norm defined in Example 1.13.

Obviously, every subspace of a normed space is a normed space by restriction Proof. We already saw that X/M is normed space. It remains to prove
of the norm. A subspace of a Banach space is a Banach space if and only if it completeness. Let ([xn ])n∈N be a Cauchy sequence.
is closed. First we show that we can assume k[xn ] − [xm ]k∼ ≤ 2−n for all m ≥
n: Choose N1 ∈ N such that k[xN1 ] − [xm ]k∼ ≤ 2−1 for all m ≥ N1 .
Proposition 1.12. Let X be a normed space. Then the following is equivalent: Next choose N2 > N1 such that k[xN2 ] − [xm ]k∼ ≤ 2−2 for all m ≥
N2 . Continuing this process, we obtain a subsequence with the desired
(i) X is complete. property. Since a Cauchy sequence converges if and only if it contains a
(ii) Every absolutely convergent series in X converges in X. convergent subsequence, it suffices to prove convergence of the subsequence
constructed above.
Proof. Exercise 1.2. By definition of the quotient norm we can assume that kxn − xn+1 k ≤
k[xn − xn+1 ]k∼ + 2−n < 21−n . Then (xn )n∈N is Cauchy sequence in X
Example 1.13 (Quotient space). Let X be a Banach space and M ⊆ X a because for all n > m
closed subspace. On X we have the equivalence relation n−1 n−1 n−1
X X X
kxn − xm k = xn+1 − xn ≤ kxn+1 − xn k < 2 2−j .
x∼y ⇐⇒ x − y ∈ M. j=m j=m j=m
14 1.2. Normed spaces Chapter 1. Banach spaces 15

Therefore x := lim xn exists and Proof. First we show that ℓp is a vector space. For α ∈ K and x, y ∈ ℓp we
n→∞
have
k[xn ] − [x]k∼ = k[xn − x]k∼ ≤ kxn − xk → 0, n → ∞. ∞ ∞
X X
|αxn |p = |α|p |xn |p < ∞
Remark 1.14. (i) In the proof above we used that, by definition of k · k∼ ,
n=1 n=1
for every x ∈ X and every ε > 0 there exists an x
e ∈ [x] such that ke
xk <
k[x]k∼ + ε. Equivalently, there exists an m ∈ M such that kx + mk < and
k[x]k∼ + ε. ∞ ∞ ∞
X X p X p
(ii) Obviously, kxk ≥ k[x]k∼ for every x ∈ X. |xn + yn |p ≤ 2 max{|xn |, |yn |} = 2p max{|xn |, |yn |}
n=1 n=1 n=1

X
Examples 1.15. (i) Finite dimensional normed spaces. Cn and Rn are com-
plete normed spaces with ≤ 2p |xn |p + |yn |p = 2p (kxkpp + kykpp ) < ∞.
n=1
k · k∞ : Kn → R, kxk∞ = max{|xj | : j = 1, . . . , n}.
Hence ℓp is a K-vector space. Properties (i) and (ii) in the definition of a
Let 1 ≤ p < ∞. Then Cn and Rn are complete normed spaces with norm are easily verified. The triangle inequality is the Minkowski inequality
(see Section 1.3).
X
n  p1
k · kp : Cn → R, kxkp = |xj |p . To show that (ℓp , k · kp ) is complete, let (xn )n∈N be a Cauchy sequence in
j=1
ℓp . Set xn = (xn,m )m∈N . Then the sequence of the m-th components is a
Cauchy sequence in K because
The triangle inequality kx+ykp ≤ kxkp +kykp is called the Minkowski inequality
(see Section 1.3). |xn,m − xk,m | < kxn − xk kp , m ∈ N.
(ii) Let T be a set and define Since K is complete, the limit ym := lim xn,m exists. Let y := (ym )m∈N .
n→∞
ℓ∞ (T ) := {x : T → K bounded map}. k·kp
We will show that y ∈ ℓp and that xn −−− → y. Let ε > 0 and N ∈ N such
Obviously, ℓ∞ (T ) is a vector space. Let that kxn − xk k < ε for all k, n ≥ N . For every M ∈ N
kxk∞ := sup{|x(t)| : t ∈ T }, x ∈ ℓ∞ , M
X
|xn,j − xk,j |p ≤ kxn − xk kpp < εp .
be the supremum norm. Then (ℓ∞ (T ), k · k∞ ) is a Banach space.
j=1

Proof. Exercise 1.4. Taking the limit k → ∞ on the left hand side yields
(iii) Sequence spaces. M
X
|xn,j − yj |p < εp .
• ℓ∞ := ℓ∞ (N) is a Banach space. j=1
• For 1 ≤ p < ∞ let
Taking the limit M → ∞ on the left hand side finally gives
n ∞
X o
ℓp := ℓp (N) := (xn )n∈N ⊆ K : |xn |p < ∞ ∞
X
n=1 |xn,j − yj |p ≤ εp < ∞,
and j=1

X
∞  p1 in particular, xn − y ∈ ℓp . Since ℓp is a vector space, we obtain y =
kxkp := |xn |p , x ∈ ℓp .
xn + (y − xn ) ∈ ℓp and kxn − ykp ≤ ε. That (xn )n∈N converges to y follows
n=1
from the inequality above since ε can be chosen arbitrarily.
With the usual component-by-component addition and multiplication with
a scalar, ℓp is a vector space and (ℓp , k · kp ) is a Banach space. (iv) Lp spaces: See measure theory.
16 1.2. Normed spaces Chapter 1. Banach spaces 17

(v) Subspaces of ℓ∞ . Let In the following, C 1 ([a, b]) will always be considered to be equipped with
the norm k · k(1) unless stated otherwise.
d := {x = (xn )n∈N ⊆ K : xn 6= 0 for at most finitely many n},
c0 := {x = (xn )n∈N ⊆ K : lim xn = 0}, Theorem 1.16. Let X be a Banach space, Y a closed subspace and N a finite
n→∞
dimensional subspace of X. Then Y + N is a closed subspace. In particular,
c := {x = (xn )n∈N ⊆ K : lim xn exists},
n→∞ every finite-dimensional subspace is closed.
Obviously, the inclusions d ( c0 ( c ( ℓ∞ hold. Moreover, it can be shown that
Proof. Obviously, Y + N is a subspace of X. To proof that it is closed, we pro-
c0 and c are closed subspaces of ℓ∞ and that d is a non-closed subspace of ℓ∞ .
ceed by induction. Therefore we can assume without restriction that dim N = 1.
In particular, (c0 , k · k∞ ) and (c, k · k∞ ) are Banach spaces, (d, k · k∞ ) is not a
Let z ∈ X such that N = {λz : λ ∈ K} and (xn )n∈N = (yn + an z)n∈N a Cauchy
Banach space (see Exercise 1.5).
sequence in Y + N .
(vi) Spaces of continuous functions. For metric space T (e. g. an interval in R) C ase 1. (an )n∈N is bounded. Then it contains a convergent subsequence
let (ank )k∈N . Then the sequence (ynk )k∈N = (xnk − ank z)k∈N converges because
it is the sum of two convergent sequences.
C(T ) := {f : T → K : f is continuous},
B(T ) := {f : T → K : f is bounded}, C ase 2. (an )n∈N is unbounded. Then there exists a subsequence (ank )k∈N with
lim |ank | = ∞. Since (xnk )k∈N is bounded, it follows that
BC(T ) := C(T ) ∩ B(T ). k→∞

1 1
For f ∈ B(T ) let
z + yn = xn → 0, n → ∞.
ank k ank k
kf k∞ := sup{|f (t)| : t ∈ T }.
Hence d(z, Y ) = 0. Since Y is closed, this implies z ∈ Y , therefore N + Y = Y
In Analysis 1 it was shown that (B(T ), k · k∞ ) and (BC(T ), k · k∞ ) are Banach
is closed in X.
spaces. Note that C(T ) = BC(T ) for a compact metric space T .
Finally, choosing Y = {0} shows that every finite-dimensional subspace is closed.
(vii) Spaces of differentiable functions. Let [a, b] a real interval. We can define
several norms on the vector space

C 1 ([a, b]) := {f : [a, b] → K : f is continuously differentiable}. Note that the sum of two closed subspaces is not necessarily closed, see as the
following example shows. Another example can be found in [Hal98, § 15].
• (C 1 ([a, b]), k · k∞ ) is not a Banach space.
1 Example. In ℓ1 consider the subspaces
Proof. For n ∈ N let fn : [−1, 1] → K, fn (t) := (t2 + n−2 ) 2 . Then
the fn converge to g : [−1, 1] → K, g(t) = |t| in the k · k∞ -norm. But U := {(xn )n∈N ∈ ℓ1 : x2n = 0, n ∈ N}
g∈/ C 1 ([a, b]). Hence C 1 ([a, b]) is not closed as a subspace of the Banach
V := {(xn )n∈N ∈ ℓ1 : x2n−1 = nx2n , n ∈ N}.
space C([a, b]), so it is not a Banach space.

• For f ∈ C 1 ([a, b]) let Obvioulsy, U and V are closed subspaces of ℓ1 . Let en be the nth unit vector
1
in ℓ1 . Let m ∈ N. Then e2m−1 ∈ U ⊆ V + U and e2m = (e2m + m e2m−1 ) −
′ 1
kf k(1) := kf k∞ + kf k∞ . m e 2m−1 ∈ V + U . Since span{en : n ∈ N} is a dense subset of ℓ 1 , it follows
1
that V + U = ℓ1 .
Then (C ([a, b]), k · k(1) ) is a Banach space. Note that the right hand side Now we will show that V + U 6= ℓ1 . Let
is finite because by assumption f ′ is continuous.
(
1
Proof. Let (xn )n∈N be a Cauchy sequence in (C 1 ([a, b]), k · k(1) ). Then x = (xn )n∈N , xn = n2 , n even,
there exist x, y ∈ C([a, b]) such that xn → x and x′n → y in the supremum 0, n odd.
norm. A well-known theorem in analysis implies x′ = y, hence xn → x in
k · k(1) . Clearly x ∈ ℓ1 . Suppose that there exist v = (vn )n∈N ∈ V , u = (un )n∈N ∈ U
18 1.2. Normed spaces Chapter 1. Banach spaces 19

such that x = v + u. It follows for all m ∈ N Theorem 1.20. All norms on Kn are equivalent.
1 Pn
= x2m = v2m + u2m = v2m , Proof. Let {e1 , . . . , en } be a basis of Kn . For x = j=1 αn en define
(2m)2
1 X
n  12
0 = x2m−1 = v2m−1 + u2m−1 = mv2m + u2m−1 = + u2m−1 , kxk2 := |αj |2 .
4m2
j=1
1
impliying that u2m−1 = − 4m 2 , m ∈ N, hence u 6= ℓ1 . Therefore x 6= V + U .
Obviously, k · k2 is a norm on X and it suffices to show thatPevery norm on X
is equivalent to k · k2 . Let k · k be a norm on X and x = nj=1 αn en . Using
Definition 1.17. Let X be a normed space and k · k1 and k · k2 be norms on
triangle inequality for k · k and Hölder’s inequality, we obtain
X. They are called equivalent norms if there exist m, M > 0 such that
Xn X n X
n  21 X
n  21

mkxk1 ≤ kxk2 ≤ M kxk1 , x ∈ X. (1.2) kxk = αj ej ≤ |αj |k ej k ≤ |αj |2 k ej k2 = C kxk2
j=1 j=1 j=1 j=1

Theorem 1.18. Let k · k1 and k · k2 be norms on a vector space X. The the (1.3)
following are equivalent: P  21
n
with constant C := j=1 k ej k2 independent of x.
(i) k · k1 and k · k2 are equivalent.
Note that k · k2 : X → R is continuous, hence S := {x ∈ X : kxk2 = 1} is
(ii) A sequence (xn )n∈N ⊆ X converges with respect to k · k1 if and only if closed being the preimage of the closed set {1} in R. In addition, S is bounded,
it converges with respect to k · k2 and in this case the k · k1 -limit and the therefore S is compact by the theorem of Heine-Borel. Now consider the map
k · k2 -limit are equal. T : (X, k · k2 ) → R, T x = kxk. By (1.3), T is uniformly continuous, so its
restriction to the compact set S has a minimum m and a maximum M . Since
(iii) A sequence (xn )n∈N ⊆ X converges to 0 with respect to k · k1 if and only k · k is a norm, m > 0 (otherwise there would exist an x ∈ S with kxk = 0, thus
if it converges with respect to k · k2 . x = 0 but 0 ∈ / S). Therefore

Proof. (i) =⇒ (ii) =⇒ (iii) is clear. mkxk2 = m ≤ kxk ≤ M = M kxk2 , x ∈ S,

“(iii) =⇒ (i)”: Obviously it suffices to show the existence of M ∈ R such and by the homogeneity of the norms
that (1.2) is true. Assume no such M exists. Then there exists a sequence
(xn )n∈N ⊆ X such that kxn k1 = 1 and kxn k2 > nkxn k1 = n. Let yn := n−1 xn . mkxk2 ≤ kxk ≤ M kxk2 , x ∈ X.
k·k1 k·k2
Then yn −−→ 0, so by assumption also yn −−→ 0. This contradicts kyn k2 > 1 The theorem above implies that all norms a a finite-dimensional K-vector space
for all n ∈ N. are equivalent. Moreover, it follows that every finite normed space is complete
because Kn with the Euclidean norm is complete and that a subset of a finite
The theorem above implies in particular, that the topologies generated by equiv- dimensional normed space is compact if and only if it is bounded and closed
alent norms coincide. Moreover, the identity map id : (X, k · k1 ) → (X, k · k2 ) is (Theorem of Heine-Borel for Kn with the Euclidean metric). In particular, the
uniformly continuous for equivalent norms. unit ball in a finite dimensional space is compact.
This is no longer true in infinite dimensional normed spaces. In fact, the unit
Example 1.19. On C 1 ([a, b]) define the norm ball is compact if and only if the dimensions of the space is finite. For the
proof we use the following theorem which is also of independent interest, as it
kf k(2) := sup{max{|x(t)|, |x′ (t)|} : t ∈ [a, b]}. shows that in a certain sense quotient spaces can work as a substitute for the
orthogonal complement in inner product spaces (see 4.2).
and let k · k(1) be as in Example 1.15 (7). It is not hard to see that
Theorem 1.21 (Riesz’s lemma). Let X be a normed space, Y ⊆ X a closed
kxk(1) ≤ kxk(2) ≤ 2kxk(1) , x ∈ C 1 ([a, b]). subspace with Y 6= X and ε > 0. Then there exists an x ∈ X such with kxk = 1
and dist(x, Y ) > 1 − ε.
20 1.2. Normed spaces Chapter 1. Banach spaces 21

Proof. If Y = {0} or ε ≥ 1, the assertion is clear. Now assume 0 < ε < 1. Note elgebraic basis) of X is a family (xλ )λ∈Λ ⊆ X that is linearly independent and
1
that in this case 1−ε > 1. Since Y is closed and different from X, the quotient such that every element x ∈ X is a (finite!) linear combination of the xλ . The
space X/Y is not trivial. Hence there exists an ξ ∈ X such that k[ξ]k∼ = 1. By existence of a Hamel basis can be shown with Zorn’s lemma.
Remark 1.14 there exists y ∈ Y such that
Definition 1.23. Let X be a normed space. A family (xn )n∈N is a Schauder
1
1 = k[ξ]k∼ ≤ kξ + yk < . basis of X if every x ∈ X can be written uniquely as
1−ε

X
Let x = kξ + yk−1 (ξ + y). Obviously, kxk = 1 and for every z ∈ Y αn xn with αn ∈ K.
n=1

kx − zk = kξ + yk−1 ξ + y − kξ + ykz ≥ kξ + yk−1 k[ξ]k∼ = kξ + yk−1 > 1 − ε.
| {z } Definition 1.24. Let (X, k · k) be a normed space over K. A subset Y ⊆ X is
∈Y
said to be a total subset of X if
Hence d(x, Y ) = inf{kx − zk : z ∈ Y } > 1 − ε.
span(Y ) = X,
Theorem 1.22. For a normed space X the following are equivalent: that is, if the set of all linear combinations of elements of Y is dense in X.
(i) dim X < ∞,
Theorem 1.25. A normed space (X, k · k) is separable if and only if it contains
(ii) BX := {x ∈ X : kxk ≤ 1} is compact. a countable total subset.

(iii) Every bounded sequence in X contains a convergent subsequence. Proof. Let A be a dense countable subset of X. Then obviously span A = X,
that is, A is a total subset of X.
Proof. “(i) =⇒ (ii)” follows from Theorem 1.20. Now assume that A is countable total subset of X. Let B := {λan : n ∈ N, λ ∈
“(ii) =⇒ (i)”: Assume that BX is compact. Then there are x1 , . . . , xn ∈ X e where Q
Q} e := Q if X is a R-vector space and Q
e := Q + iQ if X is a C-vector
with kxj k ≤ 1, j = 1, . . . , n, such that space. In both cases B is countable. We will show that B = X. Let x ∈ X
n
and ε > 0. Since A is a total subset of X, there exist a1 , . . . , an ∈ A and
[ λ1 , . . . , λn ∈ K such that
BX ⊆ B 12 (xj ). (1.4)
j=1 n
X ε
kx − λj aj k < .
Let U = span{x1 , . . . , xn }. If U 6= X, then, by Riesz’s lemma, there exists j=1
2
an x ∈ X such that kxk = 1 and dist(x, U ) > 21 , in contradiction to (1.4).
Therefore dim X = dim U ≤ n. Since Q e such that
e is dense in K, there exist µ1 , . . . , µn ∈ Q
“(ii) =⇒ (iii)”: If BX is compact, then obviously for every α ≥ 0 also αBX :=
ε X −1
n
{αx : x ∈ BX } is compact. Since every bounded sequence is a subset of some |µj − λj | < kaj k , j = 1, . . . , n.
αBX , it must contain a convergent subsequence. 2 j=1
“(iii) =⇒ (i)”: Assume that dim X = ∞. Choose x1 ∈ X with kx1 k = 1 Pn
and set U1 := span{x1 } 6= X. By Riesz’s lemma there exists an x2 ∈ X Then y := j=1 µj aj ∈ span A and
with kx2 k = 1 and dist(x2 , U1 ) > 12 , in particular kx1 − x2 k > 21 . Set U2 := Xn Xn ε X n
span{x1 , x2 } =
6 X. Continuing this way, we obtain a sequence x = (xn )n∈N ⊆ X
kx − yk ≤ x − λj aj + y − λj aj < + |µj − λj |aj
with kxn − xm k > 12 for all n, m ∈ N with n 6= m. Therefore, the sequence x j=1 j=1
2 j=1
does not contain a convergent subsequence, hence BX is not compact (Recall Xn
ε n ε ε
that a compact metric space is sequentially compact). ≤ + max |µj − λj | kan k < + = ε.
2 j=1
j=1
2 2

Let X be a vector space and Λ a set. A family (xλ )λ∈Λ ⊆ X is called linearly Note that every normed space with a Schauder basis is separable, but not every
independent if every finite subset is linearly independent. A Hamel basis (or an separable normed space has a Schauder basis.
22 1.3. Hölder and Minkowski inequality Chapter 1. Banach spaces 23

p
Examples 1.26. (i) ℓp is separable for 1 ≤ p < ∞. Theorem 1.28 (Hölder’s inequality). Let 1 ≤ p ≤ ∞ and q = p−1 , i. e.,

Proof. Let en := (0, . . . , 0, 1, 0 . . . ) be the nth unit vector in ℓp . We will show 1 1


+ =1
that {en : n ∈ N} is a total subset of ℓp . Let x = (xn )n∈N ⊆ ℓp . Then p q
n
X X
∞ (setting 1
= 0). If x ∈ ℓp and y ∈ ℓq , then z = (xn yn )n∈N ∈ ℓ1 and

x − xj ej = xj ej → 0, n → ∞.
p p
j=1 j=n+1 kzk1 ≤ kxkp kykq . (1.6)

(ii) ℓ∞ is not separable. Proof. If x = 0 or y = 0 then the inequality (1.6) clearly holds. Also the cases
p = 1 and p = ∞ are clear.
Proof. Recall that the set A := {(xn )n∈N : xn ∈ {0, 1}} i s not countable. Now assume x, y 6= 0 and 1 < p < ∞. The Young inequality (1.6) with
Obviously, A ⊆ ℓ∞ . Let B be a dense subset of ℓ∞ . Then for every x ∈ A there
exists an bx ∈ B such that kx − bx k∞ < 21 . Since kx − yk∞ = 1 for x 6= y ∈ A, it |xj | |yj |
a= , b=
follows that B has at least the cardinality of A, that is, there exists no countable kxkp kykq
dense subset of ℓp .
yields
(iii) C[a, b] is separable since by the theorem of Weierstraß the set of polyno-
mials |xj | |yj | 1 |xj |p 1 |yj |q
≤ + .
kxkp kykq p kxkpp q kykqq
{[a, b] → R, x 7→ xn : n ∈ N}
Taking the sum over gives
is a total subset of C[a, b]. ∞ ∞ ∞
1 X 1 1 X 1 1 X 1 1
|xj yj | ≤ p |xj |p + q |yj |q = + = 1.
kxkp kykq j=1 p kxkp j=1 q kykq j=1 p q
1.3 Hölder and Minkowski inequality | {z } | {z }
=kxkp
p =kykqq
In this section we prove Hölder’s inequality and Minkowski’s inequality. For the | {z } | {z }
=1 =1
proof we use Young’s inequality.
In the special case p = q = 2 we obtain the Cauchy-Schwarz inequality.
Theorem 1.27. Let p, q ∈ (1, ∞) such that
Corollary 1.29 (Cauchy-Schwarz inequality). For x = (xn )n∈N , y =
1 1
+ = 1. (yn )n∈N ∈ ℓ2 the Hölder inequality implies
p q
X∞
Then for all a, b ≥ 0:
|hx , yi| := xj yj ≤ kxk2 kyk2 .
j=1
1 p 1 q
ab ≤ a + b . (1.5)
p q
Theorem 1.30 (Minkowski inequality). For 1 ≤ p ≤ ∞ and x, y ∈ ℓp
Proof. If ab = 0, then inequality (1.5) is clear. Now assume ab > 0. Since the Minkowski’s inequality holds:
logarithm is concave and p1 + 1q = 1 is follows that
kx + ykp ≤ kxkp + kykp . (1.7)
1 1  1 1
ln ap + bq ≥ ln(ap ) + ln(bq ) = ln(a) + ln(b) = ln(ab). Proof. If x + y = 0 then (1.7) clearly holds. Also the cases p = 1 and p = ∞
p q p q
are easy to check.
Application of the monotonically increasing function exp : R → R yields (1.5). Now assume x + y 6= 0 and 1 < p < ∞. Let q ∈ (1, ∞) such that p1 + 1q = 1.
The triangle inequality in K and Hölder’s inequality (1.6) yield for all M ∈ N:
24 1.3. Hölder and Minkowski inequality Chapter 2. Bounded maps; the dual space 25

M
X M
X
|xj + yj |p = |xj + yj | · |xj + yj |p−1
j=1 j=1
M
X M
X
≤ |xj | |xj + yj |p−1 + |yj | |xj + yj |p−1
j=1 j=1 Chapter 2
p p
X
M  p1  X
M z }| {  1 X
M  p1  X
M z }| {  1
(p−1)q q (p−1)q q
≤ |xj |p |xj + yj | + |yj |p |xj + yj |


j=1 j=1
 X
M 1
q
j=1 j=1
Bounded maps; the dual
≤ kxkp + kxkp |xj + yj |p .
j=1 space
P 1
M p q
Note that j=1 |xj + yj | 6= 0 for M large enough. Hence the above
inequality yields

X
M  p1
2.1 Bounded linear maps
|xj + yj |p ≤ kxkp + kxkp
Definition 2.1. Let X, Y be normed spaces over the same field K. The set of
j=1
all linear continuous maps X → Y is denoted by L(X, Y ), i. e.,
p 1

using p − q =p 1− q = 1. Taking the limit M → ∞ finally proves (1.7).
L(X, Y ) = {T : X → Y : T linear and continuous}

and L(X) := L(X, X).

Recall that the following is equivalent:

(i) T : X → Y is continuous
(ii) lim T xn = T limn→∞ xn for every convergent sequence (xn )n∈N ∈ X
n→∞

(iii) ∀ x0 ∈ X ∀ ε > 0 ∃ δ > 0 : kx − x0 k < δ =⇒ kT x − T x0 k < ε


(iv) U ⊆ Y open =⇒ T −1 (U ) = {x ∈ X : f (x) ∈ U } open in X.

Definition 2.2. Let X, Y be normed spaces over the same field K. For a linear
map T : X → Y define the operator norm

kT k := sup{kT xk : x ∈ X, kxk = 1}.

If kT k < ∞ then T is called a bounded linear operator and kT k is the operator


norm of kT k.

Remark 2.3. (i) For a continuous linear map T : X → Y

kT xk ≤ kT k kxk, x ∈ X.
26 2.1. Bounded linear maps Chapter 2. Bounded maps; the dual space 27

Proof. The inequality is obvious for x = 0 or kxk = 1. For x ∈ X \ {0} let Theorem 2.6. Let X, Y be normed spaces.
e = kxk−1 x. By definition of kT k we find kT xk = kxk kT x
x ek ≤ kxk kT k.
Note that the inequality is also true if T is unbounded and x 6= 0. (i) L(X, Y ) is a normed space.
(ii) If Y is Banach space, then L(X, Y ) is a Banach space.
(ii) The following is easy to check:
Proof. (i) In Remark 2.4 we have seen that L(X, Y ) is a vector space. From
kT k = sup{kT xk : x ∈ X, kxk = 1} definition of the operator norm it is clear that kT k = 0 if and only if T = 0
= sup{kT xk : x ∈ X, kxk ≤ 1} and that kλT k = |λ| kT k for all λ ∈ K. To prove the triangle inequality let
n kT xk o S, T ∈ L(X, Y ) and x ∈ X.
= sup : x ∈ X, x 6= 0
kxk k(S + T )xk = kSx + T xk ≤ kSxk + kT xk ≤ kSk + kT k.
= inf{M ∈ R : ∀ x ∈ X kT xk ≤ M kxk}.
Taking the supremum over all x ∈ X with kxk = 1 yields kS + T k ≤ kSk + kT k.
Remark 2.4. (i) For S, T ∈ L(X, Y ) and λ ∈ K we define (ii) Let (Tn )n∈N be a Cauchy sequence in L(X, Y ). For x ∈ X, the sequence
(Tn x)n∈N is a Cauchy sequence in Y because
(λT + S) : X → Y, (λT + S)x := λT x + Sx.
kTn x − Tm xk ≤ kTn − Tm k kxk.
Since the sum and composition of continuous functions is continuous, and
(λT + S obviously is linear, L(X, Y ) is a vector space. Since Y is complete, we can define
It will be shown in Theorem 2.6 that k · k is indeed a norm. Note the the
T : X → Y, T x := lim Tn x.
operator norm depends on the norms on X and Y . This is can be made n→∞
explicit using the notation kT kL(X,Y ) , or similar notation.
It is easy to check that T is linear. That T is bounded and Tn → T follows as
(ii) Let X, Y, Z be normed spaces and T ∈ L(X, Y ), S ∈ L(Y, Z). Then in Example 1.13(2): For ε > 0 exists an N ∈ N such that
ε
ST : X → Z, ST x := S(T x). kTn − Tm k < , n, m ≥ N.
2
Obviously, ST ∈ L(X, Z) as composition of continuous linear functions In particular, for all x ∈ X it follows for n ≥ N that
and kST k ≤ kSk kT k because by Remark 2.3
ε
kT x − Tn xk ≤ kT x − Tm xk + kTm x − Tn xk ≤ kT x − Tm xk + , m ∈ N.
kST xk ≤ kSk kT xk ≤ kSk kT k kxk, x ∈ X. 2
(2.1)
In particular, L(X) is an algebra.
Taking the limit m → ∞ on the right hand side yields kT x − Tn xk ≤ 2ε < ε. It
Theorem 2.5. Let X, Y be normed spaces, T : X → Y linear. The following follows that T − Tn is a bounded linear map. Since L(X, Y ) is a vector space,
is equivalent: also T = Tn + (T − Tn ) is a bounded linear map. In addition, (2.1) shows that
Tn → T , n → ∞.
(i) T is continuous.
(ii) T is continuous in 0. Examples 2.7. In the following examples, the linearity of the operator under
(iii) T is bounded. consideration is easy to check.
(iv) T is uniformly continuous. (i) Let X be a normed space. Then the identity id : X → X is bounded and
k id k = 1.
Proof. The implications (iii) =⇒ (iv) =⇒ (i) =⇒ (ii) are obvious.
“ (ii) =⇒ (iii)”: Assume that T is not bounded. Then there exists a sequence (ii) Let 1 ≤ p ≤ ∞. The left shift and the right shift on ℓp are defined by
(xn )n∈N ⊆ X such that kxn k = 1 and kT xn k > n for all n ∈ N. Let yn := n−1 xn .
R : ℓp → ℓp , (x1 , x2 , x3 . . . )n∈N 7→ (0, x1 , x2 , . . . ),
Then yn → 0 but kT ynk > 1 for all n ∈ N in contradiction to the continuity of
T in 0. L : ℓp → ℓp , (x1 , x2 , x3 . . . )n∈N 7→ (x2 , x3 , . . . ).
28 2.1. Bounded linear maps Chapter 2. Bounded maps; the dual space 29

Obviously, R and L are well-defined and linear. Moreover, R is an isome- Theorem 2.9. Let X, Y be normed spaces, Y a Banach space. Let D ⊆ X be a
try because kRxkp = kxkp ; in particular kRk = 1. dense subspace of X and T ∈ L(D, Y ). Then there exists exactly one continuous
The left shift is not an isometry because, e. g., kL(1, 0, 0, . . . )kp = k0kp = extension Tb : X → Y of T . The extension is bounded with kTbk = kT k.
0 < 1 = k(1, 0, 0, . . . )kp . It is easy to see that kLxkp ≤ kxkp , x ∈ ℓp ,
implying that kLk ≤ 1. Since kL(0, 1, 0, 0 . . . )kp = k(1, 0, 0 . . . )kp = Proof. For x ∈ X choose a sequence (xn )n∈N ⊆ D which converges to x. The
k(0, 1, 0, 0 . . . )kp we also have kLk ≥ 1, so that altogether kLk = 1. sequence is a Cauchy sequence in D, hence, by the uniform continuity of T ,
Note that LR = idℓp but RL 6= idℓp . (T xn )n∈N is a Cauchy sequence in Y , and therefore it converges in Y because Y
is complete. Let (ξn )n∈N be another Cauchy sequence in D which converges to
(iii) T : C 1 ([0, 1], k · kC 1 ) → C([a, b], k · k∞ ), T x = x′ with kxkC 1 := kxk∞ + x. By what was said before, (T ξn ) converges in Y . Then lim kT xn − T ξn k =
kx′ k∞ . The operator T is bounded and kT k = 1. n→∞
lim kT (xn − ξn )k ≤ lim kT k k(xn − ξn )k = kT k lim k(xn − ξn )k = 0, the
n→∞ n→∞ n→∞
following operator is well defined:
Proof. The operator T is bounded with kT k ≤ 1 because kT xk∞ =
kx′ k∞ ≤ kxk∞ + kx′ k∞ ≤ kxkC 1 for all x ∈ X. Te : X → Y, Tex := lim T xn for any (xn )n∈N ⊆ D which converges to x.
To proof that kT k ≥ 1 let xn : [0, 1] → R, xn (t) := n1 exp(−nt). Ob- n→∞

viously, xn ∈ C 1 ([0, 1]), kxn kC 1 = n1 + 1 and kT xn k∞ = 1. It follows


It is not hard to see that Te is a linear extension of T and that kTek ≥ kT k. To
that
see that indeed equality holds, we only need to observe that by definition of Te
 xk∞  kT xn k∞
kT k = sup kT 1
kxk 1 : x ∈ C ([0, 1]) \ {0} ≥ sup kxn kC 1 : n ∈ N
 1 C {kT xk : x ∈ D, kxk = 1} = {kTexk : x ∈ X, kxk = 1},
= sup 1+ 1 : n ∈ N = 1.
n
hence the suprema of both sets without the closure are equal (and equal to
the supremum of the closed sets). Since Te is linear and bounded by kT k, it is
1 ′
(iv) T : C ([0, 1], k · k∞ ) → C([a, b], k · k∞ ), T x = x is not bounded. continuous.
Assume that S is an arbitrary continuous extension of T . For x ∈ X and a
1 sequence (xn )n∈N ⊆ D which converges to x we find
Proof. As in the example above let xn : [0, 1] → R, xn (t) := n exp(−nt).
It follows that
Sx = lim Sxn = lim T xn = lim Tbxn = Tbx.
 kT xk∞  kT xn k∞ n→∞ n→∞ n→∞
sup kxk∞ : x ∈ C 1 ([0, 1]) \ {0} ≥ sup :n∈N

kxn k∞
Therefore, Tb is the unique continuous extension of T .
= sup 11 : n ∈ N = ∞
n Finally we give a criterion for the invertibility of a bounded linear operator.

Lemma 2.8. let X, Y be normed spaces, X finite-dimensional. Then every TheoremP2.10 (Neumann series). Let X be a normed space and T ∈ L(X)

linear map T : X → Y is bounded. such that n=0 T n converges. Then id −T is invertible in L(X) and

X
Proof. Let e1 , . . . , en be a basis of X. Since on X all norms are equivalent, we
(id −T )−1 = T n. (2.2)
can assume that
n=0
Xn X n
In particular, if X is a Banach space and kT k < 1, then id −T is invertible and
αj ej = |αj |.
j=1 j=1
k(id −T )−1 k ≤ (1 − kT k)−1 .
Let M := max{kTPnej k : j = 1, . . . , n}. Then T is bounded with kT k ≤ M Proof. The proof is analogous to the proofPfor the convergence of the geometric
because for x = j=1 αj ej ∈ X m
series. We define the partial sums Sm := n=0 T n , m ∈ N0 . Then
Xn n
X n
X
(id −T )Sm = Sm (id −T ) = id −T m+1 , m ∈ N0 . (2.3)
kT xkY = αj T ej ≤ |αj | kT ej kY ≤ M |αj | = M kxkX .
Y
j=1 j=1 j=1 Note that:
30 2.1. Bounded linear maps Chapter 2. Bounded maps; the dual space 31

P∞
(i) T m → 0 for m → ∞ because m=0 T m converges. Repeating this process, it follows that
P∞
(ii) Sm → n=0 T n for m → ∞ by assumption. sn
|K n x(s)| ≤ kkkn∞ kxk∞ , s ∈ [0, 1], x ∈ C([0, 1]), n ∈ N,
(iii) For fixed R ∈ L(X) the maps L(X) → L(X), S 7→ RS and S 7→ SR n!
respectively are continuous. P∞
which shows that kK n k ≤ kkk n! . In particular,

n=0 K
n
converges so that
Hence taking the limit m → ∞ in (2.3) gives id −K is invertible by Theorem 2.10. Hence equation (2.4) has exactly one
solution x ∈ C([0, 1]), given by

X X
∞ 
(id −T ) Tn = T n (id −T ) = id ∞
X
n=0 n=0 x= K n y.
n=0
implying that id −T is invertible and that (2.2) holds. P
P ∞ P∞ P∞ kkk∞
Now assume that X is a Banach space and that kT k < 1. Then ∞ n=0 T
n Moreover, kxk∞ = n=0 K n y ≤ n=0 kK n k kyk∞ ≤ n=0 n! kyk∞ =
Pm ∞
n n
converges in norm because kT k ≤ kT k . In particular, j=0 T
j
is e kkk∞ kyk∞ .
m∈N
a Cauchy sequence in L(X). Since L(X) is complete by assumption on X
and Theorem 2.6 the series converges. By the first part of the proof, id −T is
invertible and formula (2.2) holds.
2.2 The dual space and the Hahn-Banach theo-
rem
Application 2.11 (Volterra integral equation). Let k ∈ C([0, 1]2 ) and
y ∈ C([0, 1]). We ask if the equation Definition 2.12. Let X be a normed space. X ′ := L(X, K) is the dual space
of X; elements in the dual space are called functionals.
Z s
x(s) − k(s, t)x(t) dt = y(s), s ∈ [0, 1]. (2.4)
0
Note that in general the algebraic dual space, i. e., the space of all linear maps
X → K in general is larger than the topological dual space defined above.
has solution x ∈ C([0, 1]). If a solution exists, is it unique? Can the norm of Theorem 2.6 implies immediately:
the solution be estimated in terms of y?
Proposition 2.13. The dual space of a normed space X with the norm
Solution. Note that equation (2.4) can be written as an equation in the Banach
space C([0, 1]): kx′ k = sup{|x′ (x)| : x ∈ X, kxk ≤ 1}, x′ ∈ X ′ ,

x − Kx = y is a Banach space.

where Definition 2.14. Let X be normed space. p : X → R is a a seminorm if


Z s
K : C([0, 1]) → C([0, 1]), (Kx)(s) := k(s, t)x(t) dt, s ∈ [0, 1]. (i) p(λx) = |λ|p(x), λ ∈ K, x ∈ X,
0
(ii) p(x + y) ≤ p(x) + p(y), x, y ∈ X.
Obviously, K is a well-defined linear operator and for all x ∈ C([0, 1], s ∈ [0, 1]
A seminorm p is called bounded if there exists an M ∈ R such that
Z s Z s

|Kx(s)| = k(s, t)x(t) dt ≤ |k(s, t)| |x(t)| dt ≤ s kkk∞ kxk∞ , p(x) ≤ M kxk, x ∈ X.
0 0
Z s Z t Z sZ t
If p satisfies
|K 2 x(s)| = k(s, t) k(t, t1 )x(t1 ) dt1 dt ≤ kkk2∞ kxk∞ dt1 dt
0 0 0 0
2
(i’) p(λx) = λp(x), λ ≥ 0, x ∈ X
s
= kkk2∞ kxk∞ .
2 instead of (i), then it is called a sublinear functional.
32 2.2. The dual space and the Hahn-Banach theorem Chapter 2. Bounded maps; the dual space 33

For example, every norm on X is sublinear. If ϕ : X → K is linear, then Now let c ∈ [a, b] arbitrary. We show that then ψc is an extension of ϕ0 as
desired. Let z = y + λz0 ∈ Z with y ∈ Y and λ ∈ R. Obviously ψc is continuous
X → R, x 7→ |ϕ(x)| in 0, hence ψc ∈ Z ′ .
We have to show (2.6). For λ = 0 equation (2.6) clearly holds. For λ 6= 0:
is sublinear.  
λ > 0 : λc ≤ λb ≤ λ − ϕ0 ( λ1 y) + p( λ1 y + z0 ) = −ϕ0 (y) + p(y + λz0 ),
Remark. Observe that p(x) ≥ 0 for every x ∈ X and every sublinear functional  
p. Moreover, note that every seminorm is a sublinear functional. λ < 0 : λc ≤ λa ≤ λ − ϕ0 ( λ1 y) − p( λ1 y + z0 ) = −ϕ0 (y) + p(y + λz0 ).

The next fundamental theorem shows that every normed space admits non- In both cases we obtain ψc (z) = ψc (y + λz0 ) = ϕ0 (y) + λc ≤ p(y + λz0 ) = p(z).
trivial functionals (except when X = {0}). Application to −z yields −ψc (z) = ψc (−z) ≤ p(−z) = p(z). In summary, we
have |ψ(z)| ≤ p(z), z ∈ Z.
Theorem 2.15 (Hahn-Banach theorem). Let X be normed space and p : Step 2. Let Φ be the set of all proper extensions of ϕ0 such that |ϕ(x)| ≤ p(x)
X → R a seminorm. Let Y ⊆ X a subspace and ϕ0 a linear functional on Y for all x ∈ D(ϕ) (the domain of ϕ). By Step 1, Φ is not empty and partially
(that is, ϕ0 : Y → K linear) with ordered by

ϕ0 (y) ≤ p(y), y ∈ Y. ϕ1 < ϕ2 ⇐⇒ ϕ2 is an extension of ϕ1 .

Then ϕ0 has an extension to a functional ϕ on X which satisfies Every totally ordered subset Φ0 has the upper bound
[
D(f ) = D(ψ), f (x) = ψ(x) for x ∈ D(ψ).
|ϕ(x)| ≤ p(x), x ∈ X. (2.5)
ψ∈Φ0

Proof. For Y = X there is nothing to show. Now assume Y 6= X. We dis- By Zorn’s lemma, Φ contains a maximal element ϕ. This ϕ is defined on X
tinguish between the real and the complex case. First assume that X is a real because otherwise, by Step 1, it would not be maximal.
vector space.
Now we assume that X is a complex vector space. Consider X as a vector space
We divide the proof in two steps. over R and define the functional
Step 1. Let z0 ∈ X \ Y and Z := span{z0 , Y }. We will show that ϕ0 can be
V0 : Y → R, V0 (y) = Re(ϕ(y)).
extended to some ψ ∈ Z ′ such that (2.5) holds for all z ∈ Z.
Obviously, every linear extension of ψ must be of the form It is R-linear because for all x, y ∈ Y and α ∈ R
ψc (y + λz0 ) = ϕ0 (y) + λc, λ ∈ R, y ∈ Y V0 (αx + y) = Re(ϕ0 (αx + y)) = Re(αϕ0 (x) + ϕ0 (y)) = α Re(ϕ0 (x)) + Re(ϕ0 (y))
= αV0 (x) + V0 (y).
for some c ∈ R. We have to find c such that |ψc (z)| ≤ p(z), z ∈ Z, that is,
In addition, V0 is bounded by the sublinear functional p
|ψc (y + λz0 )| ≤ p(y + λz0 ), y ∈ Y, λ ∈ R. (2.6)
|V0 (y)| = | Re(ϕ0 (y))| ≤ |ϕ0 (y)| ≤ p(y), y ∈ Y.
By assumption on ϕ0
By what we have already shown, there exists an R-linear extension V ∈ L(X, R)
ϕ0 (x) − ϕ0 (y) = ϕ0 (x − y) ≤ p(x − y) ≤ p(x + z0 ) + p(y + z0 ), y, x ∈ Y, of V0 with |V (x)| ≤ p(x), x ∈ X. Now define
ϕ : X → C, ϕ(x) = V (x) − iV (ix).
implying
ϕ has the following properties:
−ϕ0 (y) − p(y + z0 ) ≤ −ϕ0 (x) + p(x + z0 ), y, x ∈ Y,
(i) ϕ is an extension of ϕ0 . To see this, let y ∈ Y .
so that
ϕ(y) = V0 (y) − iV0 (iy) = Re(ϕ0 (y)) − i Re(ϕ0 (iy)) = Re(ϕ0 (y)) − i Re(iϕ0 (y))
a := sup{−ϕ0 (x) − p(x + z0 ) : x ∈ Y } ≤ inf{−ϕ0 (x) + p(x + z0 ) : x ∈ Y } := b. = Re(ϕ0 (y)) + i Im(ϕ0 (y)) = ϕ0 (y).
34 2.2. The dual space and the Hahn-Banach theorem Chapter 2. Bounded maps; the dual space 35

(ii) ϕ is C-linear. To show this, let x, y ∈ X and ζ = a + ib with a, b ∈ R. (i) kxk = sup{ϕ(x) : ϕ ∈ X ′ , kϕk = 1}, x ∈ X.
(ii) For T : X → Y linear
ϕ(x + y) = V (x + y) − iV (i(x + y)) = V (x) + V (y) − iV (ix) − iV (iy)
= ϕ(x) + ϕ(y), kT k = sup{ϕ(T x) : x ∈ X, kxk = 1, ϕ ∈ Y ′ , kϕk = 1}.
ϕ(ζx) = ϕ(ax) + ϕ(ibx) = V (ax) − iV (iax) + V (ibx) − iV (i2 bx)
Proof. (i) For all ϕ ∈ X ′ with kϕk = 1: kxk = kϕk kxk ≥ |ϕ(x)|, hence
= a[V (x) − iV (ix)] + b[V (ix) + iV (x)] kxk ≥ sup{ϕ(x) : ϕ ∈ X ′ , kϕk = 1}. To show that in fact we have equality, we
= (a + ib)[V (x) − iV (ix)] = ζϕ(x). recall that by Corollary 2.17 there exists a ϕ ∈ X ′ with kϕk = 1 and ϕ(x) = kxk.
Hence the formula in (i) is proved. Note the the supremum is in fact a maximum.
(iii) ϕ is bounded by p. To prove this, let x ∈ X and α ∈ R such that
   (ii) Let M := sup{ϕ(T x) : x ∈ X, kxk = 1, ϕ ∈ Y ′ , kϕk = 1}. We have to show
|ϕ(x)| = eiα ϕ(x) = Re ϕ(eiα x) = V (eiα x) ≤ p (eiα x) = p(x). M = kT k. Obviously, M = ∞ if and only if kT k = ∞. Now assume kT k < ∞.
Let ε > 0. Then there exists an x ∈ X with kxk = 1 and kT xk ≥ kT k − ε.
In conclusion, ϕ is a C-linear continuous extension of ϕ0 which is bounded by p Choose a ϕ ∈ X ′ such that kϕk = 1 and ϕ(T x) = kT xk. Then M ≥ ϕ(T x) =
as desired. kT k − ε. Since ε is arbitrary, it follows that M ≥ kT k. The revers inequality
follows from
Remark. If in the Hahn-Banach theorem we consider only real normed spaces
and replace the seminorm p by a sublinear functional such that ϕ0 (y) ≤ q(y) ϕ(T x) ≤ kϕk kT xk ≤ kϕk kT k kxk = kT k, x ∈ X, kxk = 1, ϕ ∈ X ′ , kϕk = 1.
for all y ∈ Y , then ϕ0 can be extented to a functional ϕ : X → K such that
−q(x) ≤ ϕ(x) ≤ q(x) for all x ∈ X, see [Rud91, Theorem 3.2]. Corollary 2.19. Let X be a normed space, Y ⊆ X a closed subspace. For
every x0 ∈ X \ Y exists ϕ ∈ X ′ such that ϕ|Y = 0 and ϕ(x0 ) = 1.
The Hahn-Banach theorem has some important corollaries.
Proof. Let π : X → X/Y be the canonical projection. Then π(y) = 0, y ∈ Y ,
Corollary 2.16. Let X be a normed space, Y ⊆ X a subspace and ϕ0 ∈ Y ′ . and π(x0 ) 6= 0. Since X is a normed space by Example 1.13, there exists a
Then there exists an extension ϕ ∈ X ′ of ϕ0 such that kϕk = kϕ0 k. ψ ∈ (X/Y )′ such that ϕ(π(x0 )) 6= 0 and ϕ(π(x0 )) = 1. Obviously ϕ = ψ◦π ∈ X ′
and has the desired properties.
Proof. The map p : X → R, p(x) = kϕ0 k kxk is a sublinear functional on X and
|ϕ0 (y)| ≤ kϕ0 (y)k kyk = p(y) for all y ∈ Y . By the Hahn-Banach theorem, ϕ0 Corollary 2.20. Let X be a normed space, Y ⊆ X a subspace. Then the
can be extended to a ϕ ∈ X ′ with |ϕ(x)| ≤ p(x) = kϕ0 k kxk, so that kϕk ≤ kϕ0 k. following are equivalent:
On other hand kϕk ≥ kϕ0 k holds because ϕ0 is a restriction of ϕ.
(i) Y = X,
The next corollary shows that X ′ does not consist only of the trivial functional 
(ii) ϕ|Y = 0 =⇒ ϕ = 0 , ϕ ∈ X ′.
and that it separates points in X.
Theorem 2.21. Let X be a normed space.
Corollary 2.17. Let X be a normed space, x ∈ X, x 6= 0. Then there exists a
ϕ ∈ X ′ such that ϕ(x) = kxk. In particular for all x, y ∈ X: X ′ separable =⇒ X separable.

(i) x = 0 ⇐⇒ ∀ϕ ∈ X ϕ(x) = 0,
Proof. By Proposition 1.9 the unit sphere SX ′ := {x′ ∈ X ′ : kx′ k = 1} is
(ii) x 6= y =⇒ ∃ ϕ ∈ X ′ ϕ(x) 6= ϕ(y). separable. Choose dense subset {x′n : n ∈ N} of SX ′ . and xn ∈ SX := {x ∈ X :
kxk = 1} with kx′n (xn )k > 12 . Let U = span{xn : n ∈ N}. We will show U = X.
Proof. Let Y := span{x} and ϕ0 ∈ Y ′ defined by ϕ0 (λx) = λkxk. Then ϕ0 (x) = Assume this is not true. By Corollary 2.19 there exists an x′ ∈ SX ′ such that
kxk and kϕ0 k = 1. By Corollary 2.16 there exists an extension ϕ ∈ X ′ of ϕ0 x′ 6= 0 and x′ |U = 0. Let n ∈ N such that kx′n − x′ k < 41 . This leads to the
with the desired properties. Statement (i) is clear; (ii) follows when (i) is applied contradiction
to x − y.
1 1
≤ |x′n (xn )| ≤ |x′n (xn ) − x′ (xn )| + |x′ (xn )| ≤ kx′n − x′ k + |x′ (xn )| < .
Corollary 2.18. Let X, Y be a normed spaces. 2 4
36 2.3. Examples of dual spaces Chapter 2. Bounded maps; the dual space 37

2.3 Examples of dual spaces For N large enough, the last factor in the line above is not zero, so, using
1 − 1p = q1 , we obtain
Theorem 2.22. (i) Let 1 ≤ p < ∞ and q such that
N
X  1q
1 1 |xn |q ≤ ky ′ k
+ =1
p q n=1

with the convention 1


= 0. q is called the Hölder conjugate of p. implying that x ∈ ℓq . Since (T x) en = xn en = y ′ en , n ∈ N, and {en : n ∈ N}

a total subset of ℓp , it follows that T x = y ′ . In particular, with the inequality
The following map is an isometric isomorphism:
above, kxkq ≤ ky ′ k = kT xk. Together with (2.7) it follows that kT xk = kxk,

X that is, T is an isometry.
T : ℓq → (ℓp )′ , (T x)y = xn yn for x = (xn ) ∈ ℓq , y = (yn ) ∈ ℓp . The proof for p = 1 is similar.
n=0
(ii) Well-definedness and injectivity of T are clear. Moreover kT xk ≤ kxk1 for
(ii) The following map is an isometric isomorphism: every x ∈ ℓ1 because
∞ ∞

X X X
xn yn ≤ kyk∞ |xn | = kyk∞ kxk1 , y ∈ c0 , x ∈ ℓ 1 .
T : ℓ1 → (c0 )′ , (T x)y = xn yn for x = (xn ) ∈ ℓ1 , y = (yn ) ∈ c0 .
n=0 n=0
n=0
To show that T is surjective, let y ′ ∈ (c0 )′ and let xn := yn (en ) where en is
Proof. (i) Let 1 < p < ∞. T is well-defined by Hölder’s inequality and the nth unit vector in c0 . For n ∈ N choose αn ∈ R such that |y ′ (en )| =
∞ exp(iαn )y ′ (en ). It follows that
X
|(T x)y| = xn yn ≤ kxkq kykp . ∞
X ∞
X ∞
X ∞
X 
n=0 |xn | = |y ′ (en )| = exp(iαn ) y ′ (en ) = y ′ exp(iαn ) en
n=0 n=0 n=0 n=0
Linearity and injectivity of T is clear. The inequality above gives ∞
X
kT xk ≤ kxkq , x ∈ ℓq . (2.7) ≤ ky k

exp(iαn ) en ∞ = ky ′ k.
n=0
It remains to show surjectivity of T and that kT xk ≥ kxk, x ∈ ℓq . To this end, Hence x ∈ ℓ1 and kxk1 ≤ ky ′ k. As before, since {en : n ∈ N} is a total subset
let y ′ ∈ (ℓp )′ and set xn := y ′ (en ), n ∈ N, where en is the nth unit vector in ℓp . of c0 , it follows that T x = y ′ and the proof is complete. (Note however, that
We will show that x := (xn )n∈N ∈ ℓq and that T x = y ′ . For y ′ = 0 this is clear. {en : n ∈ N} is not dense in ℓ∞ .)
Now assume that y ′ 6= 0. For n ∈ N define
( The theorem above shows that
|xn |q
xn , xn 6= 0,
tn := (ℓp )′ ∼
= ℓq , 1 ≤ p < ∞,
0, xn = 0.
(c0 )′ ∼
= ℓ1 .
Using pq − p = q we find
N N ∼ ℓ1 . Since ℓ1 is
Remark. Note that (ℓ∞ )′ ≇ ℓ1 . To see this, assume that (ℓ∞ )′ =
X X
|tn |p = |xn |p(q−1) , N ∈ N. separable, Theorem 2.21 would imply that also ℓ∞ is separable, in contradiction
n=1 n=1 to Example 1.26.
Hence, for all N ∈ N,
Other important examples are given without proof in the following theorems.
N
X N
X N
X N
X N
 X
q
|xn | = xn tn = ′
tn y (en ) = y ′
tn e n ≤ ky k

tn en p Theorem 2.23. Let (Ω, Σ, µ) be a σ-finite measure space. Let 1 ≤ p < ∞ and
n=1 n=1 n=1 n=1 n=1 q such that 1p + 1q = 1. Then
N
X  p1 N
X  p1 Z
= ky ′ k |tn |p ≤ ky ′ k |xn |q . T : Lq (Ω) → (Lp (Ω))′ , (T f )(g) = f g dµ, f ∈ Lq (Ω), g ∈ Lp (Ω),
n=1 n=1 Ω
38 2.4. The Banach space adjoint and the bidual Chapter 2. Bounded maps; the dual space 39

is an isometric isomorphism. Proof. (i) Linearity of T 7→ T ′ is clear. Immediately by the definition of T ′ we


have that
Theorem 2.24 (Riesz’s representation theorem). Let K be a compact
metric space and M (K) the set of all regular Borel measures with finite varia- kT ′ y ′ k = ky ′ ◦ T k ≤ ky ′ k kT k, y′ ∈ Y ′,
tion, that is kµk < ∞ with
hence kT ′ k ≤ kT k. By Corollary 2.18 kT k is
nX o
kµk := sup |µ(V )| : Z partition of K in pairwise disjoint measurable sets . kT k = sup{y ′ (T x) : x ∈ X, kxk = 1, y ′ ∈ Y ′ , ky ′ k = 1}.
V ∈Z
For every ε > 0 there exist x ∈ X, kxk = 1, y ′ ∈ Y ′ such that kT k − ε <
Let 1 ≤ p < ∞ and q such that 1
p + 1
q = 1. Then y ′ (T x) = (T ′ y ′ )(x) ≤ kT ′ k ky ′ k kxk = kT ′ k , so kT k ≤ kT ′k.
(ii) For all z ′ ∈ Z ′ and x ∈ X we have ((ST )′ z ′ )(x) = z ′ (ST (x)) = z ′ (S(T x)) =
Z
(S ′ z ′ )(T x) = T ′ (S ′ z ′ )x = (T ′ S ′ )(z ′ )(x), hence (ST )′ = T ′ S ′ .
T : M (K) → (C(K))′ , (T µ)(g) = g dµ, µ ∈ M (K), g ∈ C(K),

Example 2.27. Let 1 ≤ p < ∞. The adjoint of the left shift
is an isometric isomorphism.
L : ℓp → ℓp , L(x1 , x2 , x3 , . . . ) = (x2 , x3 , . . . )
For a proof, see [Rud87, Theorem 6.19]. is the right shift.
The theorems above show that
Proof. Let 1
p + 1
q = 1 and y = (yn )n∈N ∈ lq ∼
= (lp )′ . Then for all x = (xn )n∈N ∈
(Lp )′ ∼
= Lq , 1 ≤ p < ∞, lp :
(C(K))′ ∼= M (K). ∞
X ∞
X ∞
X ∞
X
(L′ y)x = y(Lx) = yn (Lx)n = yn xn+1 = yn−1 xn = (Ry)n xn
n=1 n=1 n=2 n=2
2.4 The Banach space adjoint and the bidual ∞
X
= (Ry)n xn = (Ry)x.
Definition 2.25. Let X, Y be normed spaces and T ∈ L(X, Y ). The Banach n=1
space adjoint of T is
Definition 2.28. Let X be a normed space. X ′′ := (X ′ )′ is the bidual of X.
′ ′ ′ ′ ′ ′ ′ ′
T :Y →X, (T y )x := y (T x), y ∈ Y , x ∈ X.
For every x ∈ X the linear map

Obviously, T is linear and continuous as composition of continuous functions, JX (x) : X ′ → K, JX (x)x′ := x′ x
hence T ′ ∈ L(Y ′ , X ′ ) and the following diagram commutes
is linear and bounded by kxk, hence JX (x) ∈ X ′′ .
T
X Y
Theorem 2.29. The map

x′ =y ′ ◦T y′ JX : X → X ′′ , JX (x)x′ = x′ x, x′ ∈ X ′

K is a linear isometry. In general, it is not surjective.

Theorem 2.26. Let X, Y, Z be normed spaces. Proof. We have seen above that JX is well-defined, linear and kJX (x)k ≤ kxk,
x ∈ X. Now let x ∈ X and choose ϕx ∈ X ′ such that ϕx (x) = kxk (Corol-
(i) The map L(X, Y ) → L(Y ′ , X ′ ), T 7→ T ′ , is linear and isometric, that is, lary 2.17). It follows that kJX (x)ϕx k = |ϕx (x)| = kxk, hence kJX (x)k ≥ 1.
kT ′ k = kT k. In general, it is not surjective.
The preceding theorem gives another easy proof that every normed space X can
(ii) (ST )′ = T ′ S ′ for S ∈ L(Y, Z) and T ∈ L(X, Y ). be completed (see Theorem 1.7).
40 2.4. The Banach space adjoint and the bidual Chapter 2. Bounded maps; the dual space 41

Corollary 2.30. Every normed space is isometrically isomorphic to a dense Proof. (i) Let U be a closed subspace of a reflexive normed space X and let
subspace of a Banach space. u′′ ∈ U ′′ . We have to find a u ∈ U such that JX (u) = u′′ . Let x′′0 : X ′ →
K, x′′0 (x′ ) = u′′ (x′ |U ). Obviously, x′′0 is linear and bounded because
Proof. By the theorem above, X is isometrically isomorphic to JX (X) ⊆ X ′′ .
Since X ′′ is complete (Theorem 2.6), the closure JX (X) is a Banach space. |x′′0 (x′ )| = |u′′ (x′ |U )| ≤ ku′′ kkx′ |U k ≤ ku′′ kkx′ k,

hence x′′0 ∈ X ′′ . Since X is reflexive there exists an x0 ∈ X such that JX (x0 ) =


Definition 2.31. A Banach space is called reflexive if JX is surjective. x′′0 . Assume that x0 ∈ / U . Since U is closed, there exists a ϕ ∈ X ′ such that
ϕ|U = 0 and ϕ(x0 ) = 1 (Corollary 2.19). On the other hand ϕ(x0 ) = 0 by choice
Examples 2.32. (i) Every finite-dimensional normed space is reflexive. of x0 because
(ii) ℓp is reflexive for 1 < p < ∞ by Theorem 2.22.
x′ (x0 ) = x′′0 (x′ ) = JX (x0 )x′ = u′′ (x′ |U ), x′ ∈ X ′ ,
(iii) c0 and ℓ1 are not reflexive.
Therefore x0 ∈ U . It remains to be shown that JU (x0 ) = u′′ , that is
Note that there are non-reflexive Banach spaces X such that X ∼
= X ′′ (but JX
u′′ (u′ ) = u′ (x0 ), u′ ∈ U ′ .
is not surjective).
Let u ∈ U and choose an arbitrary extension ϕ ∈ X ′ (Corollary 2.16). By
′ ′
Lemma 2.33. Let X, Y be normed spaces and T ∈ L(X, Y ). Then T ′′ ◦ JX = definition of x0 it follows that
JY ◦ T , that is, the following diagram commutes:
u′′ (u′ ) = u′′ (ϕ|U ) = x′′0 (ϕ) = ϕ(x0 ) = u′ (x0 ).
T
X Y
(ii) Let X be reflexive. We have to show that JX ′ : X ′ → X ′′′ is surjective. Let
JX JY
x′′′ ′′′ ′ ′ ′′′
0 ∈ X . The map x0 : X → K, x0 (x) = x0 (JX (x)) is linear and bounded,
hence x′0 ∈ X ′ . We will show that JX ′ (x′0 ) = x′′′ ′′ ′′
0 . Let x ∈ X . Since X is
reflexive, there exists an x ∈ X such that JX (x) = x′′ . Therefore
X ′′ Y ′′
T ′′
JX ′ (x′0 )x′′ = x′′ (x′0 ) = JX (x)(x′0 ) = x′0 x = x′′′ ′′′ ′′
0 (JX (x)) = x0 (x ),
Proof. For x ∈ X and y ′ ∈ Y ′
hence indeed JX ′ (x′0 ) = x′′′
0 .
′′ ′ ′ ′ ′ ′ ′
[T (JX (x))](y ) = (JX (x))(T y ) = T y x = y (T x) = (JY (T x))y = [(JY ◦ T )(x)]y . ′ ′ Now assume that X ′ is reflexive. By what was is already proved, X ′′ is reflexive.
Since X is a closed subspace of X ′′ via the canonical map JX , X is reflexive by
part (i) of the theorem.

If X and Y are identified with subspaces of X ′′ and Y ′′ via the canonical maps Corollary 2.36. A reflexive normed space X is separable if and only if X ′ is
JX and JY , then T ′′ is an extension of T . Note that with this identification separable.
S ∈ L(Y ′ , X ′ ) is adjoint operator of some T ∈ L(X, Y ) if and only if S ′ (X) ⊆ Y .
Proof. That separability of X ′ implies separability of X was shown in The-
Lemma 2.34. Let X be a normed space. Then ′
JX ◦ JX ′ = idX ′ . orem 2.21. If X is separable and reflexive, then also X ′′ is separable. By
Theorem 2.35 X ′ is reflexive, so we can again apply Theorem 2.21 to obtain
′ ′′′ ′ ′′′ that X ′ is separable.
Proof. Note that JX ′ : X → X and JX :X → X ′ . For x ∈ X, x′ ∈ X ′

[(JX ◦ JX ′ )x′ ](x) = [JX ′ x′ ](JX (x)) = [JX x]x′ = x′ x. Definition 2.37. Let X be a normed space. A sequence (xn )n∈N converges
weakly to x0 ∈ X if and only if
Theorem 2.35. (i) Every closed subspace of a reflexive normed space is re- lim x′ (xn ) = x′ (x0 ), x′ ∈ X ′ .
flexive. n→∞

w
(ii) A Banach space X is reflexive if and only if X ′ is reflexive. Notation: xn −
→ x or w- lim xn = x.
n→∞
42 2.4. The Banach space adjoint and the bidual Chapter 2. Bounded maps; the dual space 43

If it should be emphasised that a sequence converges with respect to the norm Continuing like this, we obtain a sequence of subsequences xnm = (xnm ,j )j∈N ,
in the given Banach space, then the sequence is called norm convergent. Some- m ∈ N such that (ϕm (xnm ,j ))j∈N converges. Now the “diagonal sequence” y
times the notion strongly convergent is used. Note, however, that in spaces with ym := xnm ,m has the desired property.
of linear operators the term “strong convergence” has another meaning (see Now we will show that y is weakly convergent. Let x′ ∈ X ′ and ε > 0. Choose
Defintion 3.12). an k ∈ N such that kx′ − ϕk k < 4M ε
where M := sup{kxn k : n ∈ N} < ∞. Let
The next remark shows that strong convergence is indeed stronger than weak N ∈ N such that |ϕk (yn ) − ϕk (ym )| < 2ε , m, n ≥ N . It follows for m, n ≥ N :
convergence.
|x′ (yn ) − x′ (ym )| ≤ |x′ (yn ) − ϕk (yn )| + |ϕk (yn ) − ϕk (ym )| + |ϕk (ym ) − x′ (ym )|
Remarks 2.38. (i) If the weak limit of a sequence exists, then it is unique, ≤ 2M kx′ − ϕk k + |ϕk (yn ) − ϕk (ym )|
because, by the Hahn-Banach theorem, the dual space separates points (Corol- ε ε
lary 2.17). < + = ε.
2 2
(ii) Every convergent sequence is weakly convergent with the same limit. This implies that (x′ (yn ))n∈N is a Cauchy sequence in K, hence it converges. To
(iii) A weakly convergent sequence is not necessarily convergent. Consider for show that (yn )n∈N converges weakly, define the map
example the sequence of the unit vectors (en )n∈N in c0 . Let ϕ ∈ c′0 ∼
= ℓ1 . Then
lim ϕ(en ) = 0 but the sequence of the unit vectors does not converge in norm. ψ : X ′ → K, ψ(x′ ) = lim x′ (yn ).
n→∞
n∈N
By what is already shown, ψ is well-defined and linear. It is also bounded
Example 2.39. Let (xn )n∈N be a bounded sequence in C([0, 1]). Then the because
following is equivalent:

|ψ(x′ )| = lim x′ (yn ) = lim |x′ (yn )| ≤ lim kx′ k k(yn )k ≤ M kx′ k.
(i) (xn )n∈N converges weakly to y ∈ C[(0, 1)]. n→∞ n→∞ n→∞

(ii) (xn )n∈N converges pointwise to y ∈ C[(0, 1)]. Hence ψ ∈ X ′′ . Since X is reflexive, there exists a y0 ∈ X such that x′ (y0 ) =
Proof. “(i) =⇒ (ii)” It is easy to see that for every t0 ∈ [0, 1] the point ψ(x′ ) = lim x′ (yn ). Hence (yn )n∈N converges weakly to y0 .
n→∞
evaluation x 7→ x(t0 ) is a bounded linear functional. Hence for all t ∈ [0, 1] the
Now assume that X is not separable. Let Y := span{xn : n ∈ N} where (xn )n∈N
sequence (xn (t)n∈N converges to some y(t). By assumption, [0, 1] → K, t 7→ y(t)
is the bounded sequence in X chosen at the beginning of the proof. Y is sep-
belongs to C([0, 1]).
arable (Theorem 1.25) and reflexive (Theorem 2.35). Hence, by the first step
“(ii) =⇒ (i)” follows from Riesz’s representation theorem (Theorem 2.24) and of the proof, there exists a subsequence (yn )n∈N ⊆ Y of (xn )n∈N and a y0 such
the Lebesgue convergence theorem (see A.19). that yn −
w
→ y0 in Y . Let x′ ∈ X ′ . Then x′ |Y ′ ∈ Y ′ , hence lim x′ (yn ) =
n→∞
w
Theorem 2.40. Every bounded sequence in a reflexive normed space contains lim x′ |Y ′ (yn ) = x′ |Y ′ (y0 ) = x′ (y0 ). Therefore we also have yn −
→ y0 in X.
n→∞
a weakly convergent subsequence.

Proof. Let X be a reflexive normed space and x = (xn )n∈N ⊆ X be a bounded


sequence.
First we assume that X is separable. By theorem 2.36, also X ′ is separable.
Let {ϕn : n ∈ N} be a dense subset of X ′ . We will construct a subsequence
y = (yn )n∈N of x such that for every j ∈ N the sequence (ϕj (yn ))n∈N converges.
The sequence (ϕ1 (xn ))n∈N is bounded, so it contains a convergent subsequence

(ϕ1 (xn1 ,1 ), ϕ1 (xn1 ,2 ), ϕ1 (xn1 ,3 ), . . . )

Now the sequence (ϕ2 (xn1 ,j ))j∈N is bounded, so it contains a convergent subse-
quence

(ϕ2 (xn2 ,1 ), ϕ2 (xn2 ,2 ), ϕ2 (xn2 ,3 ), . . . )


Chapter 3. Linear operators in Banach spaces 45

Chapter 3

Linear operators in Banach


spaces

3.1 Baire’s theorem


Theorem 3.1 (Baire-Hausdorff). Let (X, d) be a complete
T metric space and
(An )n∈N be a family of open dense subsets of X. Then ∞n=1 An is dense in X.

Taking complements, it is easily seen that the theorem above implies

Theorem. Let (X, d) be a complete


S∞ metric space and (Bn )n∈N be a family of
closed subsets of X such that n=1 Bn contains an open subset. Then at least
one of the sets Bn contains a non-empty open subset.

Proof of Theorem 3.1. For r > 0 and x ∈ X let B(x, r) := {ξ ∈ X : kx−ξk < r}.
We have to show that any open ball in X has non-empty intersection with
T
n∈N An . Let ε > 0 and x0 ∈ X.
A1 is open and dense in X, hence A1 ∩ B(x0 , ε) is open and not empty. Hence
there exist ε1 ∈ (0, 2−1 ε) and x1 ∈ A1 such that B(x1 , ε1 ) ⊆ A1 ∩ B(x0 , ε),
hence

B(x1 , ε21 ) ⊆ B(x1 , ε1 ) ⊆ A1 ∩ B(x0 , ε).

A2 is open and dense in X, hence A2 ∩ B(x1 , ε21 ) is open and not empty. Hence
there exist ε2 ∈ (0, 2−2 ε) and x2 ∈ A2 such that B(x2 , ε2 ) ⊆ A2 ∩ B(x1 , ε21 ),
hence

B(x2 , ε22 ) ⊆ B(x2 , ε2 ) ⊆ A2 ∩ B(x1 , ε21 ) ⊆ A2 ∩ A1 ∩ B(x0 , ε1 ).

In this way we obtain sequences (εn )n∈N and (xn )n∈N with 0 < εn < 2−n ε and

B(xn , ε2n ) ⊆ B(xn , εn ) ⊆ An ∩ B(xn−1 , εn−1 ) ⊆ An−1 ∩ . . . A2 ∩ A1 ∩ B(x0 , ε1 ).


(3.1)
46 3.2. Uniform boundedness principle Chapter 3. Linear operators in Banach spaces 47

Observe that xn ∈ B(xN , ε2N ) for N ∈ N and n ≥ N . This implies that (xn )n∈N Proof. For n ∈ N let
is a Cauchy sequence in X because, for fixed N ∈ N and all n, m > N we obtain \
d(xm , xn ) ≤ d(xm , xN )+d(xn , xN ) < 2−N +1 . Since X is complete, y := lim xn An := {x ∈ X : kf (x)k ≤ n}.
n→∞
f ∈F
exists and x0 ∈ B(xN , εN ) for every N ∈ N because for fixed N , we have that
xn ∈ B(xN , ε2N ) if n ≥ N . Hence (3.1) implies Note that for every n ∈ N the set {x ∈ X : kf (x)k ≤ n} is closed because f and
k · k are continuous. Since all An are intersections of closed sets, they are closed.
εN
y ∈ B(xN , ) ⊆ B(xN −1 , εN −1 ) ⊆ AN −1 ∩ . . . A2 ∩ A1 ∩ B(x0 , ε1 ),
2 N ≥ 2, Let x ∈ X. Since F is pointwise bounded, there exists an nx ∈ N such that
T x ∈ Anx , hence X ⊆ ∪n∈N An . By Baire’s theorem exists an N ∈ N, x0 ∈ X,
so y ∈ n∈N An ∩ B(x0 , ε). r > 0 such that Br (x0 ) ⊆ AN , that is, (3.2) is satisfied with M = N .

Definition 3.2. Let (X, d) be a metric space. The Banach-Steinhaus theorem is obtained in the special case of linear bounded
functions.
• A ⊆ X is called nowhere dense in X, if A does not contain an open set.
• A ⊆ X is of first category if it is the countable union of nowhere dense Theorem 3.7 (Banach-Steinhaus theorem). Let X be a Banach space, Y
sets. a normed space and F ⊆ L(X, Y ) a family of continuous linear functions which
• A ⊆ X is of second category if it is not of first category. is pointwise bounded, i. e.,

∀x ∈ X ∃ Cx ≥ 0 ∀f ∈ F kf (x)k < Cx .
Note that A is nowhere dense if and only if X \ A is dense in X.
An equivalent formulation of Then there exists an M ∈ R such that

Theorem 3.3 (Baire’s category theorem). A complete metric space is of kf k < M, f ∈ F.


second category in itself.
Proof. By the uniform boundedness principle there exists an open ball Br (x0 ) ⊆
Examples 3.4. Q is of first category in R. R is of second category in R. X and an M ′ ∈ R such that kf (x)k < M ′ for all x ∈ Br (x0 ) and f ∈ F. For
x ∈ X with kxk = 1 and f ∈ F we find
1 1
3.2 Uniform boundedness principle kf (x)k = kf (rx)k = kf (x0 ) − f (x0 − rx)k
r r
Definition 3.5. Let (X, d) be a metric space. A family F = (fλ )λ∈Λ of maps 1 2M ′
≤ (kf (x0 )k + kf (x0 − rx)k) ≤ =: M,
X → R is called uniformly bounded if there exists an M ∈ R such that r | {z } r
∈Br (x0 )

|fλ (x)| ≤ M, x ∈ X, λ ∈ Λ.
showing that F is uniformly bounded by M .
The next theorem shows that a family of pointwise bounded continuous func-
Corollary 3.8. Let X be a normed space and A ⊆ X. Then the following are
tions on a complete metric space is necessarily uniformly continuous on a certain
equivalent:
ball.
(i) A is bounded.
Theorem 3.6 (Uniform boundedness principle). Let X be a complete met- (ii) For every x′ ∈ X ′ the set {x′ (a) : a ∈ A} is bounded.
ric space, Y a normed space and F ⊆ C(X, Y ) a family of continuous functions
which is pointwise bounded, i. e., Proof. “(i) =⇒ (ii)” is clear.
“(ii) =⇒ (i)” The family (JX (a))a∈A ⊆ X ′′ is pointwise bounded by assump-
∀x ∈ X ∃ Cx ≥ 0 ∀f ∈ F kf (x)k < Cx .
tion. By the Banach-Steinhaus theorem there exists a M ∈ R such that
Then there exists an M ∈ R, x0 ∈ X and r > 0 such that kak = kJX (a)k ≤ M, a ∈ A.
∀ x ∈ Br (x0 ) ∀f ∈ F kf (x)k < M. (3.2) Hence A is bounded.
48 3.2. Uniform boundedness principle Chapter 3. Linear operators in Banach spaces 49

w
Corollary 3.9. Every weakly convergent sequence in a normed space is bounded. (iii) (Tn )n∈N converges weakly to T , denoted by w- lim Tn = T or Tn −
→ T , if
n→∞
and only if
Proof. Let X be a normed space and (xn )n∈N be a weakly convergent sequence
in X. By hypothesis, for every x′ ∈ X ′ the set {x′ (xn ) : n ∈ N} is bounded. lim |ϕ(Tn x) − ϕ(T x)| = 0, x ∈ X, ϕ ∈ Y ′ .
n→∞
Therefore, by Corollary 3.8, the set {xn : n ∈ N} is bounded.

The following theorem follows directly from Theorem 2.40 and Corollary 3.9. Remark. (i) The limits are unique if they exist.

Theorem 3.10. Let (X, k · k) be a normed space, (xn )n∈N and x0 ∈ X. Then (ii) Convergence in norm implies strong convergence and the limits are equal.
the following is equivalent: Strong convergence implies weak convergence and the limits are equal.
The reverse implications are not true:
(i) x0 = w- lim xn .
n→∞

(ii) (xn )n∈N is bounded and there exists a total subset M ′ ⊆ X ′ such that • Let X = ℓ2 (N), Tn : X → X, Tn x = (x1 , . . . , xn , 0, . . . ) for x = (xm )m∈N .
Then T converges strongly to id but kTn − id k = 1 for all n ∈ N, so that
lim f (xn ) = f (x0 ), f ∈ M ′. (Tn )n∈N does not converge to id in norm.
n→∞
• Let X = ℓ2 (N), Tn : X → X, Tn x = (0, . . . , 0, x1 , x2 , . . . ) (n leading
zeros) for x = (xm )m∈N . Then T converges weakly to 0 but kTn xk = 1 for
Corollary 3.11. Let X be Banach space and A′ ⊆ X ′ . Then the following is
all n ∈ N, so that (Tn )n∈N does not converge strongly to 0.
equivalent:

(i) A′ is bounded. Proposition 3.13. Let X be a Banach space, Y be a normed space and (Tn )n∈N ⊆
L(X, Y ) such that for all x ∈ X the limit T x := lim Tn x exists. Then T ∈
(ii) For all x ∈ X the set {a′ (x) : a′ ∈ A′ } is bounded. n∈N
L(X, Y ).
Proof. The implication “(i) =⇒ (ii)” is clear. The other direction follows di-
Proof. It is clear that T is well-defined and linear. By the uniform boundedness
rectly from the Banach-Steinhaus theorem.
principle, there exists an C ∈ R such that kTn k < C for all n ∈ N. Now let
x ∈ X with kxk = 1. Then kT xk = lim kTn xk ≤ supkTn k kxk ≤ C which
Note that for “(ii) =⇒ (i)” the assumption that X is a Banach space is necessary. n→∞ n∈N
implies that T ∈ L(X, Y ).
For example, let d = {x = (xn )n∈N : xn 6= 0 for at most finitely many n} ⊆ ℓ∞ .
d is a non-complete normed space (see Example 1.15 (5)). For m ∈ N define the
linear function ϕm : d → K by ϕm (en ) = mδm,n where δm,n is the Kronecker We finish this section with a result on strong convergence of positive operators
delta. Obviously ϕm ∈ d′ and kϕm k = m, hence the family (ϕm ) is not bounded on a space of continuous functions. An operator T on a function space is called
in d′ , but for every fixed x ∈ d the set {ϕm (x) : m ∈ M } is. positivity preserving if T f ≥ 0 for every f ≥ 0 in the domain of T .

Definition 3.12. Let X, Y be normed spaces, (Tn )n∈N ∈ L(X, Y ) a sequence Theorem 3.14 (Korovkin). Let X = C[0, 2π] the space of the continuous
of bounded linear operators and T ∈ L(X, Y ). functions on [0, 2π] and let xj ∈ X with x0 (t) = 1, x1 (t) = cos(t), x2 (t) = sin(t)
for t ∈ [0, 2π]. Let (Tn )n∈N ⊆ L(X) be a sequence of positivity preserving
(i) (Tn )n∈N converges to T , denoted by lim Tn = T , if and only if operators such that Tn xj → xj for n → ∞ and j = 0, 1, 2. Then (Tn )n∈N
n→∞
converges strongly to id, that is, Tn x → x for all x ∈ X.
lim kTn − T k = 0.
n→∞
Proof. We define the auxiliary functions
s
(ii) (Tn )n∈N converges strongly to T , denoted by s- lim Tn = T or Tn −
→ T , if t−s
n→∞ yt (s) = sin2 , t, s ∈ [0, 2π].
and only if 2

Note that yt (s) = 21 (1−cos(s) cos(t)−sin(s) sin(t)), hence yt ∈ span{x0 , x1 , x2 },


lim kTn x − T xk = 0, x ∈ X.
n→∞ in particular Tn yt → yt for n → ∞.
50 3.2. Uniform boundedness principle Chapter 3. Linear operators in Banach spaces 51

Now fix x ∈ X and ε > 0. Since x is uniformly continuous there exists a δ > 0 Note that the Fourier series is a formal series only. In the following we will
such that for all s, t ∈ [0, 2π] prove theorems on convergence of the Fourier series.
t−s First we will use methods from Analysis 1 to show that for a continuously differ-
yt (s) = sin2 <δ =⇒ |x(t) − x(s)| < ε. entiable periodic function its Fourier series converges uniformly to the function.
2
Next we will use the uniform boundedness principle to show that there exist
2kxk∞
Setting α = δ we obtain that continuous functions whose Fourier series does not converge pointwise every-
where. Finally, the Korovkin theorem implies that the arithmetic means of the
|x(t) − x(s)| ≤ ε + αyt (s), s, t ∈ [0, 2π], partial sums of the Fourier series of a periodic function converges uniformly to
because either s, t are such that yt (s) < δ, then |x(t) − x(s)| < δ by definition the function.
of δ; or yt (s) ≥ δ, then |x(t) − x(s)| ≤ 2kxk∞ = αδ ≤ αyt (s). Hence we have For a given 2π-periodic function and n ∈ N we define the nth partial sum
that n
a0 X
−ε − αyt (s) ≤ x(t) − x(s) ≤ ε + αyt (s), s, t ∈ [0, 2π] sn (x, t) = + (ak cos(kt) + bk sin(kt)). (3.3)
2
k=1
=⇒ −εx0 − αyt ≤ x(t)x0 − x ≤ εx0 + αyt , t ∈ [0, 2π]
Lemma 3.16.
and since Tn is positive and yt is a positive function ( sin((n+ 1 )s)
Z π
1 2
2 sin( 2s ) , s 6= 0,
−εTn x0 − αTn yt ≤ x(t)Tn x0 − Tn x ≤ εTn x0 + αTn yt , t ∈ [0, 2π]. sn (x, t) = x(s + t)Dn (s) ds with Dn (s) =
π −π n+ 1
s = 0.
2
Since Tn x0 → x0 and Tn yt → 21 (1 − cos(t)x1 − sin(t)x2 ) for n → ∞, we can find (3.4)
N ∈ N large enough such that εTn x0 + αTn yt < εx0 + αyt + ε for all n ≥ N ,
hence Dn is called Dirichlet kernel. Dn is continuous and
Z
|x(t)Tn x0 − Tn x| ≤ εx0 + αyt + ε, t ∈ [0, 2π], n ≥ N. 1 π
Dn (s) ds = 1. (3.5)
π −π
Hence xTn x0 − Tn x0 converges to 0 in norm in X because by the inequality
above Proof. Using the trigonometric identity cos(a) cos(b) + sin(a) sin(b) = cos(a − b)
|x(t)(Tn x0 )(t) − (Tn x)(t)| ≤ ε + αyt (t) + ε = 2ε, t ∈ [0, 2π], n ≥ N. and that x is 2π-periodic we obtain
n
That Tn x → x follows now from a0 X
sn (x, t) = + (ak cos(kt) + bk sin(kt))
2
k=1
kx − xTn x0 k∞ + kxTn x0 − Tn k∞ ≤ kxk kx0 − Tn x0 k∞ + kxTn x0 − Tn k∞ . Z π
1 1 X n 
= x(s) + (cos(ks) cos(kt) + sin(ks) sin(kt)) ds
π −π 2
k=1
Fourier Series Z 1 X n 
1 π
= x(s) + cos(k(s − t)) ds
Definition 3.15. Let x : R → R a 2π-periodic integrable function. The Fourier π −π 2
k=1
series of x is Z 1 X n 

1 π
a0 X = x(s + t) + cos(ks) ds.
S(x, t) = + (ak cos(kt) + bk sin(kt)), π −π 2
k=1
2
k=1
Now we calculate for s 6= 0
where
Z 1 X
n n
1 1 X ıs
n
1 X ıks
2n
e−ins X ıks
1 π + cos(ks) = + (e + e−iks ) = e = e
ak := x(s) cos(ks) ds, k ∈ N0 , 2 2 2 2 2
π −π k=1 k=1 k=−n k=0
Z π
1 e−ins eı2ns −1
1
1 eı(n+ 2 )s − e−ı(n+ 2 )s sin((n + 21 )s)
1
bk := x(s) sin(ks) ds, k ∈ N. = = = = Dn (s).
π −π 2 eıs −1 2 eıs/2 − e−ıs/2 2 sin 2s
52 3.2. Uniform boundedness principle Chapter 3. Linear operators in Banach spaces 53

1 1
Pn
Note that lim Dn (s) = n + 2 = 2 + k=1 cos(0). For the proof of (3.5) let Theorem 3.18. There exists a 2π-periodic continuous function x whose Fourier
s→0
x = 1 a constant function on R. Then, by (3.3), series does not converge everywhere pointwise to x.
Z
1 π Proof. We identify the 2π-periodic functions on R with
Dn (s) ds = sn (x, t) = x(t) = 1.
π −π 
X := x ∈ C([−π, π]) : x(−π) = x(π) .
Theorem 3.17. Let x : R → R be a 2π-periodic continuously differentiable Clearly (X, k · k∞ ) is a Banach space.
function. Then the Fourier series of x converges uniformly to x. Note that for fixed t ∈ [−π, π] and n ∈ N

Proof. Let x : R → R a 2π-periodic continuously differentiable function. Let sn ( · , t) : X → K


ε > 0 and h ∈ (0, π) such that h < πkxε′ k∞ . Using (3.4) and (3.5) it follows that
is linear and bounded, hence an element in X ′ .
1Z π Assume that for every x ∈ X its Fourier series converges pointwise to x. Then

|x(s) − sn (x, t)| = (x(s + t) − x(t))Dn (s) ds for every x ∈ X and t ∈ [−π, π] the sequence (sn (x, t))n∈N is bounded (because
π −π
Z it converges to x(t)). By the uniform boundedness principle there exists Ct such
1  −h Z h Z π 
that ksn ( · , t)k ≤ Ct for all n ∈ N. In particular, we have
≤ . . . ds + | . . . | ds + . . . ds .
π
| −π {z } | −h {z } | h {z } ksn ( · , 0)k ≤ C0 , n ∈ N.
=:An (t) =:Bn (t) =:Cn (t)
It is easy to see that
We have to show that An (t), Bn (t) and Cn (t) tend to 0 for n → ∞ uniformly Z Z π
in t. Using the mean value theorem and that π2 σ ≤ sin(σ) for σ ∈ [0, π/2] we 1 π
1
ksn (x, 0)k = x(s)Dn (s) ds ≤ kxk∞ |Dn (s)| ds
obtain π −π π −π
Z h Z h Rπ
|x(s + t) − x(t)| kx′ k |s| hence ksn ( · , 0)k ≤ −π |Dn (s)| ds. On the other hand, the function y(s) =
Bn (t) = s | sin((n + 12 )s)| ds ≤ s ds sign(Dn (s)) can be approximated by continuous functions ym with kym k = 1
−h 2 sin | 2 | | {z } −h 2 sin | 2 |
≤1 such that
π ε Z Z Z
≤ 2hkx′ k∞ < . 1 π 1 π 1 π
2 2 ksn (ym , 0)k = x(s)Dn (s) ds → sign(Dn (s))Dn (s) ds = |Dn (s)| ds
π −π π −π π −π
Define the auxiliary function
so that finally we obtain
x(s + t) − x(t) Z π
ft (s) = , s ∈ [h, π], t ∈ [0, π]. 1
2 sin( 2s ) ksn ( · , 0)k = |Dn (s)| ds < C0 , n ∈ N.
π −π
2kxk∞
The functions ft are continuously differentiable and kft k∞ ≤ 2 sin(h/2) =: M1 , However ksn ( · , 0)k → ∞ for n → ∞ because
kx′ k∞ Z π Z π Z π
kft′ k∞
≤ =: M2 . Note that the bounds do not depend on t. Integrating
2 sin(h/2) | sin((n + 12 )s)| | sin((n + 12 )s)|
by parts, we find |Dn (s)| ds = 2 s ds ≥ 2 ds
−π 0 2 sin 2 0 s
Z π Z π(n+ 12 )
| sin σ| X Z (k+1)π | sin σ|
n−1
Cn (t) = ft (s) sin((n + 21 )s) ds = 2 dσ ≥ 2 dσ
h σ σ
cos((n + 1 )s) π Z π cos((n + 1 )s) 0 k=0 kπ
Z (k+1)π
= − 1
2
ft (s) + 1
2
ft′ (s) ds n−1
X 1
n−1
X 1
n+ 2 h h n+ 2 ≥ 2 | sin σ| dσ = 4π .
1 M π(k + 1) kπ π(k + 1)
k=0 k=0
≤ (2M1 + (π − h)M2 ) =: . n
n + 12 n + 12 X 1
= 4 .
M ε k
Note that M ′ does not depend on t. When we choose N such that n+ 1 < 2 we
k=1
2
obtain finally |x(s) − sn (x, t)| < ε for all t ∈ R, that is, kx − sn (x, · )k∞ < ε. Hence the theorem is proved.
54 3.2. Uniform boundedness principle Chapter 3. Linear operators in Banach spaces 55

Finally we show that the arithmetic mean of the partial sums of the Fourier 3.3 The open mapping theorem
series of a continuous function converge.
Definition 3.20. A map f between metric spaces X and Y is called open if
Theorem 3.19 (Fejér). As before let the image of an open set in X is an open set in Y .

X := x ∈ C([−π, π]) : x(−π) = x(π) Note that an open map does not necessarily map closed sets to closed sets.
For example, the projection π : R × R → R, π((s, t)) = s, is open. The set
and let Tn ∈ L(X) defined by
A := {(s, t) ∈ R × R : s ≥ 0, st ≥ 2} is closed in R × R but π(A) = (0, ∞) is
1X
n−1 open in R.
Tn x = sn (x, · ).
n
k=0 Lemma 3.21. Let X, Y be Banach spaces and T ∈ L(X, Y ) such that
Then (Tn )n∈N converges strongly to id (i. e. Tn x → x for n → ∞, x ∈ X). BY (0, r) ⊆ T (BX (0, 1)).
Proof. Note that the Tn are well-defined and that for all x ∈ X and t ∈ [−π, π] for some r > 0. Then for every ε ∈ (0, 1)
n−1 Z Z π n−1
1X π 1 x(s + t)) X 1 BY (0, (1 − ε)r) ⊆ T (BX (0, 1)).
Tn x(t) = x(s + t) Dk (s) ds = s sin((k + )s) ds.
n −π nπ −π 2 sin 2 2
k=0 k=0 Here BX (x0 , r) := {x ∈ X : kx − x0 k < r} and BY (y0 , r) := {y ∈ Y : ky − y0 k <
We simplify the sum in the integrand: r} are open balls in X and Y respectively.
n−1
X n−1
X 1
 s n−1
X   s eins −1  The lemma says that if T (BX (0, 1)) is dense in BY (0, r), then, for any 0 < ρ < r,
sin((k + 21 )s) = Im ei(k+ 2 )s = Im ei 2 eiks = Im ei 2 is the ball BY (0, ρ) is contained in T (BX (0, 1)).
e −1
k=0 k=0 k=0
eins −1 eins/2 (eins/2 − eins/2 ) Proof. Note that the assertion is equivalent to
= Im = Im
eis/2 − e−is/2 eis/2 − e−is/2
BY (0, r) ⊆ (1 − ε)−1 T (BX (0, 1)) = T (BX (0, (1 − ε)−1 )).
2i(cos(ns/2) + i sin(ns/2)) sin(ns/2) sin2 (ns/2)
= Im = .
2i sin(s/2) sin(s/2) Fix ε > 0 and y0 ∈ BY (0, r). We have to show that there exists an x0 ∈ X with
kx0 k < (1 − ε)−1 and y0 = T (x0 ). By assumption, BY (0, r) ⊆ T (BX (0, 1)).
If we define the Fejér kernel Hence there exists an x1 ∈ BX (0, 1) such that ky0 − T x1 k < εr. By scaling, we
( 2
1 sin (ns/2)
s 6= 0, know that T (BX (0, ε)) is dense in BY (0, εr). Since y0 − T x1 ∈ BY (0, εr), there
2n sin(s/2) ,
Fn (s) := 1
exists an x2 ∈ BX (0, ε) such that ky0 − T x1 − T x2 k < ε2 r. Since T (BX (0, ε2 ))
2n s = 0, is dense in BY (0, ε2 r), there exists an x3 ∈ BX (0, ε2 ) such that ky0 − T x1 −
we can write Tn x as T x2 − T x3 k < ε3 r.
Z Continuing in this way, we obtain a sequence (xn )n∈N such that
1
Tn x(t) = −ππ Fn (s)x(s + t) ds. n
X
π
kxn k < εn−1 , ky0 − T xk k < rεn , n ∈ N. (3.6)
Note that all Fn are positive functions, hence the Tn are positive operators. To k=1
show the theorem, it suffices to show that Tn xj → xj for x0 (t) = 1, x1 (t) = P∞ P∞
cos(t), x2 (t) = sin(t) (Korovkin theorem). Using (3.3) it follows that sk (x0 , · ) = It
Pfollows that x0 := k=1 xk exists and lies in B(0, (1−ε)−1 ) because k=1 kxk k <
∞ k−1 −1
x0 for all k ∈ N0 and that k=1 rε = r(1 − ε) . Since T is continuous, we know that

X ∞
s0 (x1 , · ) = s0 (x2 , · ) = 0,  X
T (x0 ) = T xk = T xk .
sk (x2 , · ) = x1 , sk (x1 , · ) = x2 , k ∈ N. k=1 k=1

n−1 Pn
Since Tn x0 = x0 , Tn xj = n xj for j = 1, 2 and n ∈ N the theorem is proved. By (3.6) it follows that k=1 T xk converges to y0 for n → ∞. Hence T x0 = y0
and the statement is proved.
56 3.3. The open mapping theorem Chapter 3. Linear operators in Banach spaces 57

In the proof of the open mapping theorem we use the following fact. Proof. If rg(T ) is closed in Y then it is a Banach space. So by the previous
lemma, T : X → rg(T ) has a continuous inverse. On the other hand, if T −1 :
Remark. Let T : X → Y be a linear map between normed spaces X and Y rg(T ) → X is continuous, then T is an isomorphism between X and rg(T ), so
and assume that TX (B(0, 1)) is dense in BY (y, δ) for some y ∈ Y and δ > 0. rg(T ) is complete, hence closed in Y .
Then TX (B(0, 1)) is dense in BY (0, δ).
Corollary 3.25. Let X be a K-vector space and k · k1 and k · k2 norms on X
Proof. Obviously it suffices to show that T (BX (0, 2)) is dense in BY (0, 2δ). such that X is complete with respect to both norms. Assume that there exists
Since T is linear, it follows immediately that TX (B(0, 1)) is dense in BY (−y, δ). an α > 0 such that kxk2 ≤ αkxk1 for all x ∈ X. Then the two norms are
Let z ∈ BY (0, 2δ) and ε > 0. Note that y − z/2 ∈ BY (y, δ) and −y − z/2 ∈ equivalent.
BY (−y, δ). Choose x1 , x2 ∈ BX (0, 1) such that kT x1 − (y − z/2)k < ε/2 and
kT x2 − (−y − z)k < ε/2. Since x1 + x2 ∈ BX (0, 2) and Proof. Let T : (X, k · k1 ) → (X, k · k2 ), T x = x. T is surjective and bounded
by α, so it is continuous. By the open mapping theorem, its inverse is con-
kT (x1 + x2 ) − zk ≤ kT x1 − (y − z/2)k + kT x2 − (−y − z/2)k < ε, tinuous, hence bounded. The statement follows now from kxk1 = kT −1 xk1 ≤
kT −1k kxk2 , x ∈ X.
it follows that z ∈ T (BX (0, 2)) because ε can be chosen arbitrarily small.

Theorem 3.22 (Open mapping theorem). Let X, Y be Banach spaces and 3.4 The closed graph theorem
T ∈ L(X, Y ). Then T is open if and only if it is surjective.
Let X, Y be normed spaces. Then X × Y is a normed space with either of the
Proof. If T is open, then it is obviously surjective. norms
Now assume that T is surjective. We use the notation of the preceding lemma.
By assumption k · k : X × Y → R, k(x, y)k = kxk + kyk,
p

[ k · k : X × Y → R, k(x, y)k = kxk2 + kyk2 .
Y = T (BX (0, k)).
k=1 Note that the two norms defined above are equivalent.

Since Y is complete, by Baire’s category theorem there must exist an n ∈ N and Definition 3.26. Let X, Y be normed spaces, D a subspace of X and T : D →
y ∈ Y and ε > 0 such BY (y, ε) ⊆ T (BX (0, n)), in other words, T (BX (0, 1)) is Y linear. T is called closed if its graph
dense in BY (y/n, ε/n). By the remark above T (BX (0, 1)) is dense in BY (0, ε/n),
so by Lemma 3.21 BY (0, δ) ⊆ T (BX (0, 1)) for all δ < ε/n. G(T ) := {(x, T x) : x ∈ D} ⊆ X × Y
Now let U ⊆ X be an open set and u ∈ U . Then there exists an open ball
BX (0, ε) such that u + BX (0, ε) ⊆ U . By what was shown above, there exists is closed in X × Y . T is closable if G(T ) is the graph of an operator T . The
an δ > 0 such that T u + BY (0, δ) ⊆ T u + T (BX (0, ε)) = T (u + BX (0, ε)) ⊆ operator T is called the closure of T .
T (U ).
D is called the domain of T , also denoted by dom T . Sometimes the notations
The open mapping theorem has the following important corollaries. T : X ⊇ D → Y or T (X → Y ) are used.

Corollary 3.23 (Inverse mapping theorem). Let X, Y be Banach spaces Obviously, the graph G(T ) is a subspace of X × Y .
and T ∈ L(X, Y ) a bijection. Then T −1 exists and is continuous.
Lemma 3.27. Let X, Y normed space and D ⊆ X a subspace. Then T : X ⊇
Proof. By the open mapping theorem T is open, so its inverse T −1 is continuous. D → Y is closed if and only if for every sequence (xn )n∈N ⊆ D the following is
true:

(xn )n∈N and (T xn )n∈N converge


Corollary 3.24. Let X, Y be Banach spaces and T ∈ L(X, Y ) injective. Then (3.7)
T −1 : rg(T ) → X is continuous if and only if rg(T ) is closed. =⇒ x0 := lim xn ∈ D and lim T xn = T x0 .
n→∞ n→∞
58 3.4. The closed graph theorem Chapter 3. Linear operators in Banach spaces 59

Proof. Assume that T is closed and let (xn )n∈N such that (xn )n∈N and (T xn )n∈N Examples 3.30. (i) A continuous operator that is not closed.
converge. Then ((xn , T xn ))n∈N ⊆ G(T ) converges in X × Y . Since G(T ) is Let X be normed space, S ∈ L(X) and D a dense subset of X with
closed, lim (xn , T xn ) = (x0 , y0 ) ∈ G(T ). By definition of G(T ) this implies X \ D 6= ∅. (For example, d is dense in c0 .) Then T := S|D is continuous
n→∞
lim xn = x0 ∈ D(T ) and T x0 = y0 = lim T xn . because it is the restriction of a continuous function, but is not closed. To
n→∞ n→∞ see this, fix an x0 ∈ X \ D and choose a sequence (xn )n∈N ⊆ D which
Now assume that (3.7) holds and let ((xn , T xn ))n∈N ⊆ G(T ) be a sequence converges to x0 . Then (T xn )n∈N converges (to Sx0 ). If T were closed,
that converges in X × Y . Then both (xn )n∈N and (T xn )n∈N converge, hence this would imply that x0 ∈ D, contradicting the choice of x0 .
x0 := lim xn ∈ D and lim T xn = T x0 which shows that lim (xn , T xn ) =
n→∞ n→∞ n→∞ (ii) A closed operator that is not continuous.
(x0 , T x0 ) ∈ G(T ), hence G(T ) is closed.
Let X = C([−1, 1]), D = C 1 ([−1, 1]) ⊆ C([−1, 1]) and T : X ⊇ D →
X, T x = x′ . Then T is closed and not continuous.
Lemma 3.28. Let X, Y normed space and D ⊆ X a subspace. Then T : D → Y
is closable if and only if for every sequence (xn )n∈N ⊆ D the following is true: Proof. Let (xn )n∈N ⊆ D such that (xn )n∈N and (T xn )n∈N converge. From
a well-known theorem in Analysis 1 it follows that x0 := lim xn is differ-
lim xn = 0 and (T xn )n∈N converges =⇒ lim T xn = 0. (3.8) n→∞
n→∞ n→∞ entiable and T x0 = x′0 = ( lim xn )′ = lim x′n = lim T xn .
n→∞ n→∞ n→∞
The closure T of T is given by That T is not continuous was already shown in Example 2.7 (iv) (choose
xn (t) = n1 exp(−n(t + 1))).
D(T ) = {x ∈ X : ∃ (xn )n∈N ⊆ D with lim xn = x and (T xn )n∈N converges },
n→∞
(iii) Let X = L2 (−1, 1), D = C 1 ([0, 1]) ⊆ L2 ([0, 1]) and T : X ⊇ D →
T x = lim (T xn ) for (xn )n∈N ⊆ D with lim xn = x. X, T x = x′ . Then T is not closed.
n→∞ n→∞
(3.9) 1
Proof. Let xn : [−1, 1] → R, xn (t) = (t2 + n−2 ) 2 . Then (xn )n∈N ⊆ D
and xn → g for n → ∞ where g(t) = |t|, t ∈ [−1, 1]. The sequence of the
Proof. Assume that T is closable. Then G(T ) is the graph of a linear function.
derivatives converges
Hence for a sequence (xn )n∈N ⊆ D with lim xn = 0 and lim T xn = y for
n→∞ n→∞ 
some y ∈ Y it follows that (0, y) ∈ G(T ) = G(T ). Hence y = T 0 = 0 because T 
1, t > 0,
′ t
is linear. xn (t) = 1 → h(t) = −1, t < 0,
(t2 + n−1 ) 2 

Now assume that (3.8) holds and define T as in (3.9). T is well-defined because 0, t = 0.
for sequences (xn )n∈N and (x̃n )n∈N in D with lim xn = lim x̃n = x such that
n→∞ n→∞ Obviously h ∈ L2 (−1, 1). If T were closed, it would follow that g ∈
(T xn )n∈N and (T x̃n )n∈N in D converge, it follows that (xn − x̃n )n∈N converges C 1 ([−1, 1]), a contradiction.
to 0. Since T (xn − x̃n ) = T (xn − x̃n ) converges, it follows by assumption that
lim T xn − lim T x̃n = lim T (xn − x̃n ) = 0. Linearity of T is clear. By Definition 3.31. Let X, Y be Banach spaces, D ⊆ X a subspace and T : X ⊇
n→∞ n→∞ n→∞
definition, G(T ) is the closure of G(T ), so T is the closure of T . D → Y a linear operator. Then
k · kT : D → R, kxkT = kxk + kT xk
Remarks 3.29. Let X, Y be normed spaces.
is called the graph norm of T .
(i) Every T ∈ L(X, Y ) is closed.
It is easy to see that k · kT is a norm
p on D. Moreover, the norm defined above
(ii) If T is closed and injective, then T −1 is closed.
is equivalent to the norm kxk′T = kxk2 + kT xk2 on D. Most of the time, the
graph norm defined in Definition 3.31 is easier to use in calculations. However,
Proof. Closedness of {(x, T x) : x ∈ X} ⊆ X × Y implies closeness of the norm with the square root is sometimes more useful when operators in
{(T −1 y, y) : y ∈ rg(T )} ⊆ X × Y . Hilbert spaces are considered.

(iii) If T : D ⊇ X → Y is linear and continuous, then T is closable and Lemma 3.32. Let X, Y be Banach spaces, D ⊆ X a subspace and T : X ⊇
D(T ) = D(T ). D → Y a closed linear operator. Then
60 3.4. The closed graph theorem Chapter 3. Linear operators in Banach spaces 61

(i) (D, k · kT ) is a Banach space. Lemma 3.36. Let X, Y be Banach spaces, D ⊆ X a subspace and T : D → Y
linear. Then the following are equivalent:
(ii) Te : (D, k · kT ) → Y, Tex = T x, is continuous.
(i) T is closed and D(T ) is closed.
Proof. (i) To show completeness of (D, k · kT ) let (xn )n∈N ⊆ D be a Cauchy (ii) T is closed and T is continuous.
sequence with respect to k · kT . Then, by definition of the graph norm, (xn )n∈N (iii) D(T ) is closed and T is continuous.
is a Cauchy sequence in X and (T xn )n∈N is a Cauchy sequence in Y . Since
X and Y are complete, the sequences converge. Hence, by the closeness of T , Proof. (i) =⇒ (ii) follows from the closed graph theorem because by assumption
k·kT
k · k- lim xn =: x0 ∈ D and xn −−−→ x0 . D is Banach space.
n→∞
(ii) =⇒ (iii) and (iii) =⇒ (i) are clear.
(ii) The statement follows from kTexkY ≤ kxkX + kT xkY = kxkT , x ∈ D.
Example 3.37. An everywhere defined linear operator that is not closed.
Lemma 3.33. Let X, Y be Banach spaces, D ⊆ X a subspace and T : X ⊇
D → Y a closed surjective operator. Then T is open. If, in addition, T is
Let X be an infinite dimensional Banach space and (xλ )λ∈Λ an algebraic basis
injective, then T −1 is continuous.
of X. Without restriction we can assume kxλ k = 1, λ ∈ Λ. Choose N → Λ, n 7→
λn be an injection. Then the operator
Proof. By Lemma 3.32 and the open mapping theorem (Theorem 3.22) the
X X
operator iTe : (D, k · kT ) → Y, Tex = T x, is open. Let U ⊆ D open with respect T : X → X, T (x) = n cλn xλn for x = cλn xλn ∈ X,
to the norm in X. Then U is also open with respect to the graph norm because n∈N λ∈Λ
obviously i : (D, k · kT ) → (D, k · k), ix = x, is bounded, hence continuous.
Hence T (U ) = Te(U ) is open in Y . is well-defined. Assume that T is closed. By the closed graph theorem T must
be bounded, but kT xλn k = knxλn k = n while kxλn k = 1, n ∈ N contradicting
Now assume in addition that T is injective. Then Te−1 : Y → (D, k · kT ) is
the boundedness of T .
continuous by the inverse mapping theorem. Since i is continuous, also T −1 =
(Te ◦ i−1 )−1 = i ◦ Te−1 is continuous.
3.5 Projections in Banach spaces
Lemma 3.34. Let X, Y be Banach spaces, D ⊆ X a subspace and T : X ⊇ D →
Y a closed injective linear operator such that T −1 : rg(T ) → X is continuous. Definition 3.38. Let X be a vector space. P : X → X is called a projection
Then rg(T ) is closed. (on rg(P )) if P 2 = P .

Proof. Let (yn )n∈N be a Cauchy sequence in rg(T ) with y0 := lim yn . and Note that if P is a projection, then also id −P is a projection because (id −P )2 =
n→∞
xn := T −1 yn , n ∈ N. Then (xn )n∈N is a Cauchy sequence in D because kxn − id −2P + P 2 = id −P .
xm k = kT −1 yn − T −1 ym k ≤ kT −1k kyn − ym k. Hence (xn )n∈N converges in X
and its limit x0 belongs to D and y0 = lim yn = T x0 ∈ rg(T ) because T is Lemma 3.39. Let X be a normed space and P ∈ L(X) a projection. Then the
n→∞
following holds:
closed.
(i) Either P = 0 or kP k ≥ 1.
Theorem 3.35 (Closed graph theorem). Let X, Y be Banach spaces and
T : X → Y be a closed linear operator. Then T is bounded. (ii) ker(P ) and rg(P ) are closed.
(iii) X is isomorphic to ker P ⊕ rg(P ).
Proof. Note that the projections
Proof. (i) Note that kP k = kP 2 k ≤ kP k2 , hence 0 ≤ kP k − kP k2 = kP k(1 −
π1 : G(T ) → X, π1 (x, T x) = x,
kP k).
π2 : G(T ) → Y, π2 (x, T x) = T x
(ii) Since P is continuous, ker(P ) = P −1 ({0}) is closed. To see that rg(P ) is
are continuous and that π1 is bijective. By assumption the graph G(T ) is closed closed, it suffices to show that rg(P ) = ker(id −P ). Indeed, x ∈ ker(id −P )
in X × Y , hence a Banach space, so π1 is open by the open mapping theorem implies x = P x ∈ rg(P ) and y ∈ rg(P ) implies (P − id)y = P y − y = y − y = 0,
(Theorem 3.22). Hence T = π2 ◦ π1−1 is continuous. hence y ∈ ker(id −P ).
62 3.6. Weak convergence Chapter 3. Linear operators in Banach spaces 63

(iii) Obviously x 7→ ((id −P )x, P x) ∈ ker(P ) ⊕ rg(P ) is well defined, linear, generated by U, denoted by τ (U).
bijective and continuous because id −P and P are continuous. By the inverse
mapping theorem then also the inverse operator is continuous which shows that Obviously τ (U) exists and is the intersection of topologies containing all Uλ .
X and ker(P ) ⊕ rg(P ) are isomorphic.
Lemma 3.45. Let X be a set, U = (Uλ )λ∈Λ a family of subsets of X. Then
Theorem 3.40. Let X be a normed space, U ⊆ X a finite dimensional sub- the topology generated by U consists of all sets of the form
space. Then there exists a linear continuous projection P of X to U with n
[ \
kP k ≤ dim U . Uγ,k , (3.10)
γ∈Γ k=1
Proof. From linear algebra we know that there exist bases (u1 , . . . , un ) of U and that is, of arbitrary unions of finite intersections of sets in the family U.
(ϕ1 , . . . , ϕn ) of U ′ such that kuk k = kϕk k = 1 and ϕj (uk ) = δjk , j, k = 1, . . . , n.
By the Hahn-Banach theorem the ϕk can be extended to linear functionals ψk Proof. Let τ (U) be the topology generated by U and σ(U) the system of sets
on X with kϕk k = kψk k. We define described in (3.10). It is not hard to see that σ(U) is a topology containing U,
n
X hence containing τ (U). On the other hand, all sets of the form (3.10) are open
P : X → X, Px = ϕk (x)uk . in τ (U), so σ(U) ⊆ τ (U).
k=1
Pn
Obviously P is a linear bounded projection on U and kP xk ≤ kϕk k kxk kuk k = Definition 3.46. Let X be a set, Λ be an index set and for every λ ∈ Λ let
Pn k=1
(Yλ , τλ ) be a topological space. Consider a family F = (fλ : X → Yλ ) of
k=1 kxk = nkxk.
functions. The smallest topology on X such that all fλ are continuous, is called
Theorem 3.41. Let X be Banach space, U, V ⊆ X closed subspaces such that the initial topology on X, denoted by σ(X, F ).
X and U ⊕ V are algebraically isomorphic. Then the following holds: 
(i) X is isomorphic to V ⊕ U with k(u, v)k = kuk + kvk. Note that τ (F ) = τ {fλ−1 (Uλ ) : λ ∈ Λ, Uλ ∈ τλ }.
(ii) There exists a continuous linear projection of X on U .
(iii) V is isomorphic to X/U . Definition 3.47. Let X be a normed space. The topology σ(X, X ′ ) is called
the weak topology on X. The topology σ(X ′ , X) is called the weak ∗ topology on
Proof. (i) Since U and V are Banach spaces, their sum U ⊕ V is a Banach space. X ′ when X is identified with a subset of X ′′ by the canonical map JX .
The map U ⊕ V → X, (u, v) 7→ u + v is linear, continuous and bijective. Hence
by the inverse mapping theorem, also the inverse is continuous. Note that σ(X ′ , X) ⊆ σ(X ′ , X ′′ ) ⊆ σk·k .
(ii) P : X → U, u + v 7→ u is the desired projection. Lemma 3.48. Let X be a normed space. A sequence (xn )n∈N ⊆ X is weakly
(iii) The map V 7→ X/V, v 7→ [v] is linear, bijective and continuous. Since U is convergent to some x0 ∈ X (in the sense of Definition 2.37) if and only if it
closed, X/U is a Banach space. By the inverse mapping theorem it follows that converges in the weak topology σ(X, X ′ ).
V and X/U are isomorphic.
Proof. Assume that (xn )n∈N is weakly convergent with x0 := w- lim xn and let
Definition 3.42. let X be a Banach space. A closed subspace U i of X is called n→∞
complemented if there exists a continuous linear projection on U . U be a σ(X, X ′ )-open set containing x0 . Then there exist ϕ1 , . . . , ϕn such that
n
\
Remark 3.43. Note that not every closed subspace of a Banach space is com- x0 ∈ ϕ−1
j (Vj ) ⊆ U
plemented in the sense of the theorem above. For example, c0 is not comple- k=1

mented as subspace of ℓ∞ . with Vj open subsets in R containing ϕj (x0 ). Since lim ϕ(xn ) = ϕ(x0 ) for all
n→∞
ϕ ∈ X ′ , we can choose an TN ∈ N such that ϕj (xn ) ∈ Uj for all n ≥ N and all
n
3.6 Weak convergence j = 1, . . . , n. Hence xn ∈ k=1 {ϕ−1
j (Vj )} ⊆ U for all n ≥ N .
Now assume that (xn )n∈N ⊆ X converges to x0 in the weak topology. Since
Definition 3.44. Let X be a set and U = (Uλ )λ∈Λ a family of subsets of sets in by definition of σ(X, X ′ ) all functionals ϕ ∈ X ′ are continuous, it follows that
X. The smallest topology on X such that all Uλ are open is called the topology (ϕ(xn ))n∈N converges to ϕ(x0 ) for every ϕ ∈ X ′ .
64 3.6. Weak convergence Chapter 3. Linear operators in Banach spaces 65

Lemma 3.49. Let X be a normed space, (xn )n∈N ⊆ X and (ϕn )n∈N ⊆ X ′ . Theorem 3.53 (Banach-Alaoglu). Let X be a normed space. Then the closed
unit ball K1′ := {ϕ ∈ X ′ : kϕk ≤ 1} is weak ∗-compact.
(i) x0 = w- lim xn =⇒ kx0 k ≤ lim inf xn .
n→∞ n→∞
Proof. For x ∈ X define the set Ax := {z ∈ K : |z| ≤ kxk} and let A :=
Q
(ii) ϕ0 = w ∗ - lim ϕn =⇒ kϕ0 k ≤ lim inf ϕn .
n→∞ n→∞ x∈X Ax together with the product topology. By Tychonoff’s theorem A is
compact. Note that elements a ∈ A are maps X → K with |a(x)| ≤ kxk, x ∈ X.
Proof. (i) For x0 = 0 the assertion is clear. By the Hahn-Banach theorem there Hence K1′ ⊆ A because |ϕ(x)| ≤ kϕk kxk ≤ kxk for every ϕ ∈ K1′ . The product
exists an ϕ ∈ X ′ such that ϕ(x0 ) = kx0 k and kϕk = 1. Hence topology on A is the weakest topology on A such that for every x ∈ X the map
πx : A → K, a 7→ a(x) is continuous. Hence the topology on K1′ induced by A
kx0 k = k lim ϕ(xn )k ≤ lim inf kϕk kxn k = lim inf kxn k. is exactly the weak ∗-topology on K1′ . So it suffices to show that K1′ is closed in
n→∞ n→∞ n→∞
A with the product topology.
(ii) Let ε > 0. Then there exists an x ∈ X with kxk = 1 such that kϕ0 k − ε < Let ϕ ∈ K1′ and let x, y ∈ X and ε > 0. Then
kϕ0 (x)k. The statement follows as above:
U := {a ∈ A : |a(x + y) − ϕ(x + y)| < ε, |a(x) − ϕ(x)| < ε, |a(y) − ϕ(y)| < ε}
kϕ0 k − ε < kϕ0 (x)k = lim kϕn (x)k ≤ lim inf kϕn k kxk = lim inf kϕn k.
n→∞ n→∞ n→∞
is an open neighbourhood of ϕ. Hence there exists an g ∈ K1′ ∈ U ∩ K1′ . Since
Definition 3.50. Let X be a topological space. A function f : X → R is called g is linear, it follows that
upper semicontinuous if lim sup f (xn ) ≤ f (x). It is called lower semicontinuous
xn →x
|ϕ(x + y) − ϕ(x) − ϕ(y)| = |ϕ(x + y) − ϕ(x) − ϕ(y) − g(x + y) + g(x) + g(y)|
if lim inf f (xn ) ≥ f (x).
xn →x
≤ |ϕ(x + y) − g(x + y)| + |ϕ(x) − g(x)| + |ϕ(y) − g(y)| < 3ε.
Hence the lemma above states that k · k is lower semicontinuous in the weak Since ε was arbitrary, this implies ϕ(x + y) = ϕ(x) + ϕ(y). Similarly it can be
topology. shown that ϕ(λx) = λϕ(x) for λ ∈ K and x ∈ X. It follows that ϕ is linear.
Since ϕ ∈ A, it follows that kϕk ≤ 1, hence ϕ ∈ K1′ .
Definition 3.51. For λ ∈ Λ let (Xλ , τλ ) be topological spaces. Define
Y n [ o
X := Xλ := f : Λ → Xλ : f (λ) ∈ Xλ , λ ∈ Λ .
λ∈Λ λ∈Λ

The product topology on X is the weakest topology such that for every λ ∈ Λ
the projection

πλ : X → Xλ , πj (f ) = f (j),

is continuous.

Lemma 3.52. Let X as above with the product topology. Let O ⊆ P(X) be the
family of all sets U ⊆ X such that for every u ∈ U there exist λj ∈ Λ, Uj ⊆ Xλj
open, j = 1, . . . , n, such that
n
\
u ∈ {s ∈ X : s(λj ) ∈ Uj , j = 1, . . . , n} = πλ−1
j
(Uj ) ⊆ U.
j=1 | {z }
open in O

Then O is the product topology on X.

Proof. This is a special case of Lemma 3.48.


Chapter 4. Hilbert spaces 67

Chapter 4

Hilbert spaces

4.1 Hilbert spaces


Definition 4.1. Let X be a K-vector space. A map

h· , ·i : X × X → K

is a sesquilinear form on X if for all x, y, z ∈ X, λ ∈ K

(i) hλx + y , zi = λhx , zi + hy , zi,


(ii) hx , λy + zi = λhx , yi + hx , zi.

The inner product is called

• hermitian ⇐⇒ hx , yi = hy , xi, x, z ∈ X,
• positive semidefinite ⇐⇒ hx , xi ≥ 0, x ∈ X,
• positive (definite) ⇐⇒ hx , xi > 0, x ∈ X \ {0}.

Definition 4.2. A positive definite hermitian sesquilinear form on a K-vector


X is called an inner product on X and (X, h· , ·i) is called an inner product space
(or pre-Hilbert space).

Note that hx , xi ∈ R, x ∈ X, for a hermitian sesquilinear form X because


hx , xi = hx , xi.

Lemma 4.3 (Cauchy-Schwarz inequality). Let X be a K-vector space with


inner product h· , ·i. Then for all x, y ∈ X

|hx , yi|2 ≤ |hx , xi| |hy , yi|, (4.1)

with equality if and only if x and y are linearly dependent.


68 4.1. Hilbert spaces Chapter 4. Hilbert spaces 69

Proof. For x = 0 or y = 0 there is nothing to show. Now assume that y 6= 0. Theorem 4.7 (Polarisation formula). Let X be an inner product space over
For all λ ∈ K K and x, y ∈ X. Then

0 ≤ hx + λy , x + λyi = hx , xi + λhy , xi + λhx , yi + |λ|2 hy , yi. 1 


hx , yi = kx + yk2 − kx − yk2 , if K = R,
4
In particular, when we choose λ = − hx ,yi
hy ,yi we obtain
1 
hx , yi = kx + yk2 − kx − yk2 + ikx + iyk2 − ikx − iyk2 , if K = C.
4
|hy , xi|2 |hx , yi|2 |hx , yi|2
0 ≤ hx + λy , x + λyi = hx , xi − − + Proof. Straightforward calculation.
hy , yi hy , yi hy , yi
|hx , yi|2
= hx , xi −
hy , yi A necessary and sufficient criterion for a normed space to be an inner product
space is the following.
which proves (4.1). If there exist α, β ∈ K such that αx+βy = 0, then obviously
equality holds in (4.1). On the other hand, if equality holds, then hx + λy , x +
Theorem 4.8 (Parallelogram identity). Let X be normed space. Then the
λyi = 0 with λ chosen as above, so x and y are linearly dependent.
norm on X is generated by an inner product if and only if for all x, y ∈ X the
parallelogram identity is satisfied:
Note that (4.1) is true also in a space X with a semidefinite hermitian sesquilin-
ear form but equality in (4.1) does not imply that x and y are linearly dependent.
kx + yk2 + kx − yk2 = 2kxk2 + 2kyk2.
Lemma 4.4. An inner product space (X, h· , ·i) becomes a normed space by
1
setting kxk := hx , xi 2 , x ∈ X. In this case, the inner product is given by the polarisation formula.

Proof. The only property of a norm that does not follow immediately from the Proof. Assume that the norm is generated by the inner product h· , ·i and let
1
definition of k · k is the triangle inequality. To prove the triangle inequality, kxk = hx , xi 2 . Then for all x, y ∈ X parallelogram identity holds:
choose x, y ∈ X. Using the Cauchy-Schwarz inequality, we find
kx + yk2 + kx − yk2 = kxk2 + kyk2 + 2 Rehx , yi + kxk2 + kyk2 − 2 Rehx , yi
kx + yk2 = kxk2 + 2 Rehx , yi + kyk2 ≤ kxk2 + 2|hx , yi| + kyk2
= 2kxk2 + 2kyk2 .
≤ kxk2 + 2kxk kyk + kyk2 = (kxk + kyk)2 .
Now assume that the norm on X is such that the parallelogram identity holds
In the following, we will always consider inner product spaces endowed with the
and for x, y ∈ X define hx , yi by the polarisation formula. We prove that h· , ·i
topology induced by the norm.
is an inner product on X in the case K = C. The case K = R can be proved
analogously.
Definition 4.5. A complete inner product space is called a Hilbert space.

Lemma 4.6. Note that the scalar product on a inner product space X is a • Positivity.
continuous map X × X → K when X × X is equipped with the norm k(x, y)k =
kxkX + kykX . 4hx , xi = kx + xk2 − kx − xk2 + i kx + ixk2 − i kx − ixk2
= 4kxk2 + i kx + ixk2 − i kix + xk2 = 4kxk2 ≥ 0.
Proof. The statement follows from

|hx1 , x2 i − hy1 , y2 i| = |hx1 , x2 − y2 i − hy1 − x1 , y2 i| • Hermiticity.


≤ kx1 k kx2 − y2 k − ky1 − x1 k ky2 k.
4hx , yi = kx + yk2 − kx − yk2 + i kx + iyk2 − i kx − iyk2
The polarisation formula allows to express the inner product of two elements of
X in terms of their norms. = ky + xk2 − ky − xk2 + i k − ix + yk2 − i kix + yk2 = 4hy , xi.
70 4.1. Hilbert spaces Chapter 4. Hilbert spaces 71

• Additivity. (ii) ℓ2 (N) with



X
4(hx , yi + hx , zi) hx , yi = xk yk , x = (xk )k∈N , y = (yk )k∈N ,
= kx + yk2 − kx − yk2 + i kx + iyk2 − i kx − iyk2 k=1

+ kx + zk2 − kx − zk2 + i kx + izk2 − i kx − izk2 is an inner product space.


y + z y − z y + z y − z
2 2
= x + + − x − − (iii) Let R([0, 1]) be the vector space of the Riemann integrable functions on
2 2 2 2
y+z y − z y+z y − z the interval [0, 1]. Then
2 2
+ x + − − x − +
2 2 2 2 Z 1
− 2 y −z
y + z y z y + z 2 hf , gi = f (t)g(t) dt, f, g ∈ R([0, 1]),
+ i x + i +i − i x − i −i
2 2 2 2 0
y − z y −z
y+z 2 y+z 2
+ i x + i −i − i x − i +i defines a sesquilinear form on R([0, 1]) which is not positive definite, since,
2 2 2 2 for example, χ{0} 6= 0, but hχ{0} , χ{0} i = 0.
y + z y − z 2 y + z y − z 2
2 2
= 2 x + + 2 − 2 x − − 2 The restriction of h· , ·i to the space of the continuous functions C([0, 1]) is
2 2 2 2
y + z 2 y − z 2 y + z 2 y − z 2 an inner product which is not complete (its closure is the space L2 ([0, 1])).

+ 2i x + i + 2i − 2i x − i − 2i
2 2 2 2
2 2 2 y + z
y + z y + z y +z 2
= 2 x +
2
− 2 x −
2
+ 2i x + i
2
− 2i x − i
2
4.2 Orthogonality
y+z
= 2 · 4hx , i. Definition 4.11. Let X be an inner product space.
2
(i) Elements x, y ∈ X are called orthogonal, denoted by x ⊥ y, if and only if
If we choose z = 0 we find hx , yi = 2hx , y2 i, hence
hx , yi = 0

y + z (ii) Subsets A, B ⊆ X are called orthogonal, denoted by A ⊥ B, if and only if
hx , yi + hx , zi = 2 x , = hx , y + zi.
2 ha , bi = 0 for all a ∈ A, b ∈ B.
(iii) The orthogonal complement of a set M ⊆ X is
• Homogeneity. From the additivity we obtain hλx , yi = λhx , yi for all λ ∈
Q. Note that hix , yi = ihx , yi, hence homogeneity is proved for λ ∈ Q+iQ. M ⊥ := {x ∈ X : x ⊥ m, m ∈ M }.
Hence for fixed x, y ∈ C the two continuous functions C → C, λ 7→ λhx , yi
and C → C, λ 7→ hλx , yi must be equal because they are equal on the Remarks 4.12. (i) Pythagoras’ theorem holds: kx + yk2 = kxk2 + kyk2 if
dense subset Q + iQ of C. x ⊥ y.
(ii) For every set M ⊆ X its orthogonal complement M ⊥ is a closed subspace
Theorem 4.9. The completion of an inner product space is an inner product
of X.
space.
(iii) A ⊆ (A⊥ )⊥ for every subset A ⊆ X.
Proof. By continuity of the norm, the parallelogram identity holds on the com- (iv) A⊥ = (span A)⊥ for every subset A ⊆ X.
pletion X of an inner product space X. So X is an inner product space.
Theorem 4.13 (Projection theorem). Let H be a Hilbert space, M ⊆ H a
Examples 4.10. (i) Rn and Cn with the Euclidean inner product nonempty closed and convex subset and x0 ∈ H. Then there exists exactly one
y0 ∈ M such that kx0 − y0 k = dist(x0 , M ).
n
X
hx , yi = xk yk , x = (xk )nk=1 , y = (yk )nk=1 , Proof. Recall that dist(x0 , M ) := inf{kx0 − yk : y ∈ M }. If x0 ∈ M then the
k=1
assertion is clear (choose y0 = x0 ).
are inner product spaces. Now assume that x0 ∈
/ M . Without restriction we may assume x0 = 0.
72 4.2. Orthogonality Chapter 4. Hilbert spaces 73

Existence of y0 . Let d := dist(x0 , M ) = inf{kyk : y ∈ M }. Then there exists Proof. (i) =⇒ (ii) Let y ∈ U . If y = 0, then obviously hx0 − y0 , yi = 0. If
a sequence (yn )n∈N ⊆ M such that lim kyn k = d. We will show that (yn )n∈N kyk = 1, let λ = kyk−1 hx0 − y0 , yi. By assumption
n→∞
is a Cauchy sequence. Note that ≥ d2 because yn +y
k yn +y
2
m 2
k 2
m
∈ M by the kx0 − y0 k2 ≤ kx0 − y0 − λyk2
convexity of M . Hence the parallelogram identity (Theorem 4.8) yields
= kx0 − y0 k2 − λhx0 − y0 , yi − λhy , x0 − y0 i + |λ|2 kyk2
y − y 2 y − y 2 y + y 2
n m n m n m = kx0 − y0 k2 + (1 − 2kyk−2)|hx0 − y0 , yi|2
≤ + − d2
2 2 2
1 = kx0 − y0 k2 − |hx0 − y0 , yi|2
= (kyn k2 + kym k2 ) − d2 −→ 0, n, m → ∞.
2 so hx0 − y0 , yi = 0. By linearity of U then x0 − y0 ⊥ y for all y ∈ U .
Since X is a Banach space, (yn )n∈N converges to some y0 ∈ X, and since M is (ii) =⇒ (i) Let y ∈ U . By assumption
closed, y0 ∈ M .
kx0 − yk2 = k(x0 − y0 ) + (y0 − y)k2 = kx0 − y0 k2 + ky0 − yk2 ≥ kx0 − y0 k2 .
Uniqueness of y0 . Assume that there are y0 , ỹ0 ∈ M such that ky0 k = kỹ0 k =
d = dist(x0 , M ). The parallelogram identity yields Recall that a linear operator P : X → X on a Banach space X is called a
y + ỹ 2 y + ỹ 2 y − ỹ 2 projection if and only if P 2 = P (see Definition 3.38).
0 0 0 0 0 0 1
d2 ≤ ≤ + = (ky0 k2 + kỹ0 k2 ) = d2 .
2 2 2 2 Theorem 4.16. Let H be a Hilbert space, U ⊆ H a closed subspace with U 6=
{0}. Then there exists a projection PU ∈ L(H) on U such that kPU k = 1 and
It follows that ky0 − ỹ0 k = 0, so y0 = ỹ0 .
ker(PU ) = U ⊥ . Also id −PU is continuous projection with k id −PU k = 0 if
U = H and k id −PU k = 1 if U 6= H. If U ⊕ U ⊥ is equipped with the norm
Lemma 4.14. Let M be a closed and convex subset of a Hilbert space H and 1
k(u, v)k = (kuk2 + kvk2 ) 2 , then H = U ⊕ U ⊥ .
fix x0 ∈ H. For y0 ∈ M the following are equivalent:
(i) kx0 − y0 k = dist(x0 , M ), Definition 4.17. PU as in the theorem is called the orthogonal projection on
(ii) Rehx0 − y0 , y − y0 i ≤ 0, y ∈ M. U.

Proof. (i) =⇒ (ii) For t ∈ [0, 1] and y ∈ M let yt := y0 + t(y − y0 ). Then Proof of Theorem 4.16. Fix x0 ∈ H and let PU (x0 ) := y0 the unique element
yt ∈ M by the convexity of M and by assumption on y0 y0 ∈ U such that kx0 − y0 k = dist(x0 , U ). Then rg(PU ) = U and PU2 = PU ,
hence PU is a projection on U .
kx0 − y0 k2 ≤ kx0 − yt k2 = kx0 − y0 − t(y − y0 )k2 By Lemma 4.15, PU (x0 ) is the unique element in U such that x0 −PU (x0 ) ∈ U ⊥ .
= kx0 − y0 k2 − 2t Rehx0 − y0 , y − y0 i + t2 ky − y0 k2 . Rehx0 − PU (x0 ) , y − PU (x0 )i ≤ 0, y ∈ U.
So for all t ∈ (0, 1] We will show that PU is linear. Let x1 , x2 ∈ H and λ ∈ K. Since U ⊥ is a
subspace, we obtain
2 Rehx0 − y0 , y − y0 i ≤ tky − y0 k2
λx1 − x2 − (λPU (x1 ) − PU (x2 )) = λ(x1 − PU (x1 )) − (x2 − PU (x2 )) ∈ U ⊥ .
which implies Rehx0 − y0 , y − y0 i ≤ 0.
Hence, by definition of PU ,
(ii) =⇒ (i) Let y ∈ M . By assumption
PU (λx1 − x2 ) = λPU (x1 ) − PU (x2 ).
kx0 − yk2 = k(x0 − y0 ) + (y0 − y)k2
= kx0 − y0 k2 + ky0 − yk2 + 2 Rehx0 − y0 , y0 − yi ≥ kx0 − y0 k2 . We already know that rg(PU ) = U . ker(PU ) = U ⊥ because
PU (x) = 0 ⇐⇒ x0 ∈ U ⊥ .
Lemma 4.15. Let U be a closed subspace of a Hilbert space H and fix x0 ∈ H.
For y0 ∈ U the following are equivalent: Therefore id −PU is a projection with rg(id −PU ) = U ⊥ and ker(id −U ) = U .
By Pythagoras’ theorem we obtain
(i) kx0 − y0 k = dist(x0 , U ),
(ii) x0 − y0 ⊥ U . kx0 k2 = kPU (x0 ) + (id −PU )(x0 )k2 = kPU (x0 )k2 + k(id −PU )(x0 )k2 .
74 4.2. Orthogonality Chapter 4. Hilbert spaces 75

In particular, H = U ⊕ U ⊥ with norm as in the statement, and kPU k ≤ 1 and Proof. (ii) is clear. Let Ψ : H ′ → H ′′ as in Theorem 4.20. Then it is easy to
k id −PU k ≤ 1. Lemma 3.39 implies kPU k = 1, k id −PU k = 1 if U 6= H and check that Ψ ◦ Φ = JH , so JH is surjective, implying that H is reflexive.
k id −PU k = 0 if U = H.
Corollary 4.22. Let H be a Hilbert space.
Lemma 4.18. Let U be a subspace of a Hilbert space H. Then U = U ⊥⊥ .
(i) A sequence (xn )n∈N ⊆ H converges weakly to x0 ∈ H if and only if
Proof. By the projection theorem (Theorem 4.16), for every closed subspace V
hxn − x0 , yi → 0, y ∈ H.
PV = id −PV ⊥ = id −(id −PV ⊥⊥ ) = PV ⊥⊥ ,
(ii) Every bounded sequence (xn )n∈N ⊆ H contains a weakly convergent sub-
hence V = V ⊥⊥ . Application to V = U shows the statement. sequence.

Definition 4.19. Let X, Y be vector spaces. A map X → Y is called antilinear Proof. (i) follows from the Riesz-Fréchet theorem, and (ii) follows with Theo-
or conjugate linear if f (λx + y) = λf (x) + f (y) for all λ ∈ K and x, y ∈ X. rem 2.40.

Theorem 4.20 (Fréchet-Riesz representation theorem). Let H be a Hilbert


space. Then the map 4.3 Orthonormal systems
Φ : H → H ′, y 7→ h· , yi Definition 4.23. Let H be a Hilbert space. A family S = (xλ )λ∈Λ of vectors
in H is called an orthonormal system if hxλ , xλ′ i = δλλ′ . A orthonormal system
is an isometric antilinear bijection. S is an orthonormal basis (or a complete orthonormal system) if and only if for
every orthonormal system T
Proof. Obviously Φ(0) = 0 ∈ H ′ . The Cauchy-Schwarz inequality yields
S ⊆ T =⇒ S = T.
kΦ(y)(x)k = |hx , yi| ≤ kxk kyk, x, y ∈ H,
Examples 4.24. (i) The unit vectors (en )n∈N in ℓ2 (N) are a orthonormal
hence kΦ(y)k ≤ kyk If y 6= 0, then set x = kyk−1 y. Note that kxk = 1 and system.
kΦ(y)xk = kyk, implying that kΦ(y)k = kyk. So we have shown that Φ is
(ii) Let H = L2 (−π, π). An orthonormal system in H is
well-defined and an isometry. In particular, Φ is injective.
To show that Φ is surjective, fix an ϕ ∈ H ′ . If ϕ = 0, then ϕ = Φ(0). Otherwise n 1 o n 1 o n 1 o
we can assume that kϕk = 1. Since ker{ϕ} is closed, there exists a decomposition S= √ ∪ √ sin(n · ) : n ∈ N ∪ √ cos(n · ) : n ∈ N .
2π π π
H = ker ϕ ⊕ (ker ϕ)⊥ . Note that rg(ϕ) = K, hence dim(ker ϕ)⊥ = 1. Choose
y0 ∈ (ker ϕ)⊥ with ϕ(y0 ) = 1. Then (ker ϕ)⊥ = span{y0 }. For x = u + λy0 ∈ Lemma 4.25 (Gram-Schmidt). Let H be a Hilbert space and (xn )n∈N a
ker ϕ ⊕ (ker ϕ)⊥ , family of linearly independent vectors. Then there exists a orthonormal system
S = (sn )n∈N such that span S = span{xn : n ∈ N}.
hx , ky0 k−2 y0 i = λ = λϕ(y) + ϕ(u) = ϕ(x),
Proof. Let s1 := kx1 k−1 x1 . Next set y2 := x2 − hx1 , s1 is1 . Note that y2 6= 0
hence ϕ = h· , ky0 k−1 y0 i. Since Φ is an isometry, it follows that 1 = kϕk =

ky0 k because x2 and x1 are linearly independent. Let s2 := ky2 k−1 y2 . Then s1 ⊥ s2
ky0 k2 = ky10 k , so ky0 k = 1. and ks1 k = ks2 k = 1. Now for k ≥ 1 let
n
X
Corollary 4.21. (i) Every Hilbert space is reflexive.
yn+1 := xn+1 − hxk , sk isk , sn+1 := kyn+1 k−1 yn+1 .
(ii) The dual H ′ of a Hilbert space H is an inner product space by k=1

Since x1 , . . . , xn+1 are linearly independent, sn+1 is well-defined. By con-


hΦ(x) , Φ(y)iH ′ = hy , xiH
struction, sn+1 ⊥ sj for j = 1, . . . n. Note that for every n ∈ N, sn ∈
with Φ : H → H ′ as in Theorem 4.20. span{x1 , . . . , xn } and xn ∈ span S, hence span S = {xn : n ∈ N}.
76 4.3. Orthonormal systems Chapter 4. Hilbert spaces 77

Example. Let H = L2 ((0, 1)) and xn ∈ H defined by xn (t) = tn .qApplication Proof. For fixed x ∈ H, the set Sx = {s ∈ S : hx , si =
6 0} is at most countable
of the Gram-Schmidt orthogonalisation yields polynomials sn (t) = n + 21 Pn (t) (Lemma 4.27), so the claim follows from the Bessel inequality for countable
1 dn 2
orthonormal systems.
where Pn (t) = 2n n! dtn (t − 1)n is the nth Legendre polynomial.
Theorem 4.30. Let H be a Hilbert space and S ⊆ H a orthonormal system.
Theorem 4.26 (Bessel inequality). Let H be a Hilbert space, {sn : n ∈ N} Then
a orthonormal system in H. Then X
∞ P : H → H, Px = hx , sis
X
|hx , sn i|2 ≤ kxk2 , x ∈ H. s∈S

n=1
is an orthogonal projection on span S and the series is unconditionally conver-
PN gent.
Proof. For N ∈ N let xN := x− n=1 hx , sn isn . Since xN ⊥ sn for n = 1, . . . , N ,
Pythagoras’ theorem yields
Proof. First we proof that the series in the definition of P is unconditionally
XN 2 N
X N
X convergent (this proves then well-definedness of P ). Fix x ∈ H. For fixed
2 2
kxk = kxN k + hx , sn isn = kxN k2 + |hx , sn i|2 ≥ |hx , sn i|2 . x ∈ H, the set Sx = {s ∈ S : hx , si =
6 0} is at most countable
Pn (Lemma 4.27).
n=1 n=1 n=1 Let Sx = {sn : n ∈ N} be an enumeration of Sx . Then k=1 hx , sk isk n∈N is
a Cauchy sequence because

XM 2 M
X

Lemma 4.27. Let H be a Hilbert space, S = (sλ )λ∈Λ a orthonormal system in hx , sk isk = |hx , sk i|2 −→ 0, M, K → ∞
H. Then for every x ∈ H the set k=N k=N
P∞
Sx := {λ ∈ Λ : hx , sλ i =
6 0} by Bessel’s inequality. Since H is complete, yP:= k=1 hx , sk isk exists. Let
π : N → N be a permutation. Then also yπ := ∞ k=1 hx , sπ (k)isπ(k) exists. We
is at most countable. have to show that y = yπ . For all z ∈ H

X ∞
X
Proof. By the Bessel inequality, for every n ∈ N the set
hy , zi = hy , sn ihsn , zi = hy , sπ ihsπ , zi = hyπ , zi.
n 1o n=1 n=1
Sx,n := λ ∈ Λ : |hx , sλ i| ≥
n P
S∞ We have used that ∞ n=1 hy , sn ihsn , zi is absolute convergent and can therefore
is finite. Hence Sx = n=1 Sx,n is at most countable. be rearranged, because, by Hölder’s inequality and Bessel’s inequality
P X
∞ 2  X
∞  X
∞ 
Definition 4.28. Let X be a normed space, (xλ )λ∈Λ ⊆ X. Then λ∈Λ xλ
converges unconditionally to x ∈ H if and only if Λ := {λ ∈ Λ : x 6
= 0} is at |hy , sn ihsn , zi| ≤ |hy , sn i|2 |hsn , zi|2 ≤ kyk2kzk2 < ∞.
P 0 λ
n=1 n=1 n=1
most countable and ∞ n=1 xλn = x for every enumeration Λ0 = {λn : n ∈ N}.
Since y − yπ ⊥ z, z ∈ H, it follows that y = yπ . Therefore the series in the
Recall that in finite dimensional Banach spaces unconditional convergence is definition of P is unconditionally convergent and P is well-defined.
equivalent to absolute convergence. In every infinite dimensional Banach space, It is clear that P is a linear and kP k ≤ 1 follows from Corollary 4.29. Let
however, there exists a unconditionally convergent series that does not converge ⊥
absolutely (Dvoretzky-Rogers theorem). x ∈ H. We have to show that x − P x ∈ span S (Theorem 4.16). This is clear
because
D X E D X E
Corollary 4.29 (Bessel inequality). Let H be a Hilbert space and S ⊆ H a x− hx , si , s0 = x − hx , si , s0 = 0, s0 ∈ S.
orthonormal system. Then s∈S s∈Sx
X
|hx , si| ≤ kxk2 , x ∈ H. Theorem 4.31. Let H be a Hilbert space and S ⊆ H a orthonormal system.
s∈S Then the following is equivalent.
78 4.3. Orthonormal systems Chapter 4. Hilbert spaces 79

(i) S is a complete orthonormal system. (iii) Let H = L2 (0, 1) and


(ii) x ⊥ S =⇒ x = 0, x ∈ H. n 1 o n 1 o n 1 o
S= √ ∪ √ sin(n · ) : n ∈ N ∪ √ cos(n · ) : n ∈ N .
(iii) H = span S. 2π π π
X Note that span S is the set of all trigonometric polynomials. Without
(iv) x = hx , sis, x ∈ H.
s∈S
restriction we can assume that K = R. By the theorem of Fejér, the
X trigonometric polynomials are dense in C2π := {f ∈ C([−π, π]) : f (−π) =
(v) hx , yi = hx , sihs , yi, x, y ∈ H. f (π)} with respect to k · k∞ , hence also with respect to k · k2 . Since C2π
s∈S is k · k2 -dense in L2 ([−π, π]), S is a total subset of L2 ([−π, π]).
X
(vi) Parseval’s equality holds: kxk2 = |hx , si|2 , x ∈ H. Lemma 4.33. Let H be an infinite dimensional Hilbert space. Then the fol-
s∈S
lowing is equivalent.
Proof. (i) =⇒ (ii) If there exists an x ∈ H such that x ∈ S ⊥ \ {0}, then (i) H is separable.
S ′ := S ∪ {kxk−1 x} is a orthonormal system with S ( S ′ , contradicting the (ii) Every complete orthonormal system in H is countable.
maximality of S.
(iii) There exists an countable complete orthonormal system in H.
(ii) =⇒ (iii) follows from Lemma 4.18.
P Proof. (i) =⇒ (ii) Assume S ⊆ H is an uncountable complete orthonormal
(iii) =⇒ (iv) By theorem 4.30, x 7→ s∈S hx , sis is the orthogonal projection 1

on span S = H. system in H. Let ε ∈ (0, 2−p2 ) and s 6= s′ ∈ S. ′


√ Then Bε (s) ∩ Bε (s ) = ∅ because

by Pythagoras ks − s k = ksk2 + ks′ k2 = 2. Let A be a dense subset of H.
(iv) =⇒ (v) straightforward. For every s ∈ S there exists an as ∈ A such that as ∈ Bε (s). In particular,
(v) =⇒ (vi) Choose x = y. as 6= as′ if s 6= s′ , so A cannot be countable, thus H is not separable.
(vi) =⇒ (i) Assume there exists an orthonormal system S ′ ) S. Then for (ii) =⇒ (iii) The existence of a complete orthonormal system in H follows
every s′ ∈ S ′ \ S we get the contradiction from Zorn’s lemma. By assumption, it must be complete.
X (iii) =⇒ (i) Let S be a countable orthonormal system in H. Then span S = H
1 = ks′ k2 = |hs′ , si|2 = 0. by Theorem 4.31 and H is separable by Theorem 1.25.
s∈S
Lemma 4.34. Let H be Hilbert space and S and T be complete orthonormal
Now we show that the orthonormal systems in Example 4.24 are complete. system in H. Then |S| = |T |.

Examples 4.32. (i) The set of the unit vectors {en : n ∈ N} in ℓ2 (N) are a Proof. The statement is proved in linear algebra if |S| < ∞. Now assume that
complete orthonormal system in ℓ2 (N) because {en : n ∈ N} = ℓ2 (N). S is not finite. For x ∈ S the set Tx := {y ∈ TS : hx , yi =
6 0} is at most countable
by Lemma 4.27. By Theorem 4.31 (ii) T ⊆ x∈S Tx , hence |T | ≤ |S||N| = |S|.
(ii) Let Γ be a set and define Analogously, |S| ≤ |T ||N| = |T |. By the Schröder-Bernstein theorem then
n |S| = |T |.
ℓ2 (Γ) := f : Γ → K : f (γ) 6= 0 for at most countably many γ ∈ Γ and
X o Theorem 4.35. Let H be a Hilbert space and S an orthonormal basis of H.
|f (γ)|2 < ∞ . Then H ∼
= ℓ2 (S) (see Example 4.32 (ii)).
γ∈Γ

P Proof. Define T : H → ℓ2 (S) by T x(s) = hx , si, x ∈ H, s ∈ S. T is well-defined


Then hf , gi = γ∈Γ f (γ)g(γ) is a well-defined inner product (note that by Bessel’s inequality. Then T : H → ℓ2 (S) is linear and isometric by
only countably many terms are 6= 0 and the sum is absolutely convergent PParseval’s
equality. To show that T is surjective, let y ∈ ℓ2 (S) and define x := s∈S y(s)s.
by Hölder’s inequality). As in the case Γ = N it can be shown that ℓ2 (Γ) Then x ∈ H (Theorem 4.30) and T x = y.
is a Hilbert space and (fλ )λ∈Γ where fλ (γ) = δλγ (Kronecker delta) is a
complete orthonormal system in ℓ2 (Γ). Note that by construction hT x , T yi = hx , yi, x, y ∈ H.
80 4.4. Linear operators in Hilbert spaces Chapter 4. Hilbert spaces 81

Corollary 4.36. If H is a separable Hilbert space, then H ∼


= ℓ2 (N). Remarks. (i) T selfadjoint =⇒ T normal.

Corollary 4.37 (Fischer-Riesz theorem). L2 [0, 1] ∼


= ℓ2 (N). (ii) T ∈ L(H1 , H2 ) =⇒ T T ∗ and T ∗ T are selfadjoint.

Next we show that a length preserving linear map between Hilbert spaces also
4.4 Linear operators in Hilbert spaces preserves angles.

Definition 4.38. Let H1 , H2 be Hilbert spaces and Φj : Hj → Hj′ the canon- Lemma 4.41. Let H1 , H2 be Hilbert spaces and T ∈ L(H1 , H2 ).
ical isomorphism in the Fréchet-Riesz representation theorem (Theorem 4.20).
Let T ∈ L(H1 , H2 ). Its (Hilbert space) adjoint operator is T ∗ := Φ−1 ′ (i) T is an isometry ⇐⇒ hT x , T yi = hx , yi, x, y ∈ H1 .
1 T Φ2 ∈
L(H2 , H1 ) where T ′ is the Banach space adjoint of T (see Definition 2.25). (ii) T is unitary ⇐⇒ T is a surjective isometry.

Hence T ∗ is characterised by Proof. (i) The direction “⇐” is clear; “⇒” follows from the polarisation formula
(Theorem 4.7).
hT x , yi = hx , T ∗ yi, x ∈ H1 , y ∈ H 2 . (ii) “⇒” Since T is unitary, if follows that rg(T ) ⊇ rg(T T ∗) = rg(idH2 ) = H2 ,
so T is surjective. T is an isometry because for all x, y ∈ H1
Theorem 4.39. Let H1 , H2 , H3 be Hilbert spaces, S, T ∈ L(H1 , H2 ), R ∈
L(H2 , H3 ) and λ ∈ K. hT x , T yi = hT ∗ T x , yi = hx , yi,

(i) (λS + T )∗ = λS ∗ + T ∗ .
(ii) (RT )∗ = T ∗ R∗ . “⇐” Assume that T as a surjective isometry. Since
(iii) T ∗ ∈ L(H2 , H1 ) and kT ∗k = kT k.
hx , y − T ∗ T yi = hx , yi − hT x , T yi = 0, x, y ∈ H1 ,
(iv) T ∗∗ = T .
(v) kT T ∗k = kT ∗ T k = kT k2 . ∗ ∗
it follows that T T y = y, so T T = idH1 . In particular T ∗ is surjective. Now
(vi) ker T = (rg(T ∗ ))⊥ , ker T ∗ = (rg(T ))⊥ . we will show that T ∗ is an isometry. Let ξ, η ∈ H2 . Then there exist x, y ∈ H1
(vii) If T is invertible, then (T −1 )∗ = (T ∗ )−1 . such that T x = ξ and T y = η. It follows that

hT ∗ ξ , T ∗ ηi = hT ∗ T x , T ∗ T yi = hx , yi = hT x , T yi = hξ , ηi.
Proof. (i)–(iv) are clear. For the proof of (v) note that for kxk = 1
By the same argument as for T we conclude that idH2 = T ∗∗ T ∗ = T T ∗ .
kT xk2 = hT x , T xi = hx , T ∗ T xi ≤ kxk kT ∗T xk ≤ kT ∗ T k ≤ kT ∗ k kT k = kT k2.
Examples 4.42. (i) Let H1 , H2 be Hilbert spaces with dim H1 = dim H2 =
Taking the supremum over all x ∈ H with kxk = 1 shows the desired equalities.
n < ∞. After choice of bases, a linear operator T : H1 → H2 has a
representation (aij )nij=1 ∈ Mn (C). The matrix corresponding to T ∗ is
(vi) ker T = (rg T ∗ )⊥ because for x ∈ H then (aji )nij=1 .

Tx = 0 ⇐⇒ ∀y ∈ H2 hT x , yi = 0 ⇐⇒ ∀y ∈ H2 hx , T ∗ yi = 0 (ii) Let H = L2 [0, 1]. For k ∈ L∞ ([0, 1] × [0, 1]) define


⇐⇒ x ⊥ rg(T ∗ ). Z 1
Tk : L2 [0, 1] → L2 [0, 1], (Tk f )(t) = k(s, t)f (s) ds.
Then also ker T ∗ = (rg(T ∗∗ ))⊥ = (rg(T ))⊥ . 0

Definition 4.40. Let H1 , H2 be Hilbert spaces, T ∈ L(H1 , H2 ). Then Tk ∈ L2 [0, 1] and


Z 1
(i) T is called unitary if T is invertible and T T ∗ = idH2 and T ∗ T = idH1 . Tk∗ : L2 [0, 1] → L2 [0, 1], (Tk f )(t) = k(s, t)f (s) ds,
∗ ∗ 0
(ii) T is called normal if H1 = H2 and T T = T T .
(iii) T is called selfadjoint if H1 = H2 and T = T ∗ . that is Tk∗ = Tk .
82 4.4. Linear operators in Hilbert spaces Chapter 4. Hilbert spaces 83

Theorem 4.43 (Hellinger-Toeplitz). Let H be a Hilbert space, T : H → H Hence, by the parallelogram identity (Theorem 4.8), for kxk ≤ 1, kyk ≤ 1,
a linear operator such that
1 
hT x , yi = hx , T yi, x, y ∈ H. RehT x , yi ≤ |hT (x + y) , x + yi| + |hT (x − y) , x − yi|
4
1  M
Then T is bounded, hence selfadjoint. ≤ M kx + yk2 + M kx − yk2 = (kxk2 + kyk2 ) ≤ M.
4 2
Proof. It suffices to show that T is closed because D(T ) = H is closed. Let Now choose λ ∈ C, |λ| = 1 such that λhT x , yi = |hT x , yi|, so
(xn )n∈N ⊆ H with xn → 0 and T xn → y. Observe that
|hT x , yi| = hT (λx) , yi = | RehT (λx) , yi| ≤ M, kxk ≤ 1, kyk ≤ 1.
kyk2 = lim hT xn , yi = lim hxn , T yi = h lim xn , T yi = h0 , T yi = 0,
n→∞ n→∞ n→∞
In particular, kh· , T xik ≤ M , so kT xk ≤ 1 for kxk ≤ 1. This shows kT k ≤
so y = 0. This implies that T is closable, hence closed since D(T ) = H.
M.
Theorem 4.44. Let H be a complex Hilbert space. For T ∈ L(H) the following
Corollary 4.46. Let H be a Hilbert space and T ∈ L(H) selfadjoint. If
is equivalent.
hT x , xi = 0, x ∈ H, then T = 0.
(i) hT x , xi ∈ R, x ∈ H.
(ii) T is selfadjoint. Note that the condition hT x , xi = 0 automatically implies that T is selfadjoint
in the case of a complex Hilbert space. In a real Hilbert spaces H the assumption
Proof. (ii) =⇒ (i) follows from that T is selfadjoint is necessary for the statement in the corollary. For example,
 0 1
let T = : R2 → R2 the rotation about 90◦ . Then T 6= 0 but
hT x , xi = hx , T xi = hT x , xi, x ∈ H. −1 0
hT x , xi = 0 for all x ∈ R2 .
(i) =⇒ (ii) Let x, y ∈ H and λ ∈ C.

A := hT (λx + y) , λx + yi = |λ|2 hT x , xi + hT y , yi + λhT x , yi + λhT y , xi, Lemma 4.47. Let H be a Hilbert space, T ∈ L(H) a normal operator. Then

B := hT (λx + y) , λx + yi = |λ|2 hT x , xi + hT y , yi + λhy , T xi + λhx , T yi. kT xk = kT ∗ xk, x ∈ H,

By assumption, A = B, so in the special cases λ = 1 and λ = i we obtain in particular, ker T = ker T ∗ .


hT x , yi + hT y , xi = hy , T xi + hx , T yi,
Proof. 0 = hT ∗ T x − T T ∗ x , xi = kT xk2 − kT ∗ xk2 .
hT x , yi − hT y , xi = −hy , T xi + hx , T yi,

so finally hT x , yi = hx , T yi. Definition 4.48. Let H be a Hilbert space. A bounded selfadjoint operator
T ∈ L(H) is called non-negative, denoted by T ≥ 0, if hT x , xi ≥ 0 for all x ∈ H.
Theorem 4.45. Let H be a Hilbert space, T ∈ L(H) selfadjoint. Then It is called positive, denoted by T > 0, if hT x , xi > 0 for all x ∈ H \ {0}. We
write T ≤ S if and only if S − T ≥ 0. A sequence (Tn )n∈N ∈ L(H) is increasing
kT k = sup |hT x , xi|. if and only if Tn ≤ Tn+1 , n ∈ N. A sequence (Tn )n∈N ∈ L(H) is decreasing if
kxk≤1
and only if (−Tn )n∈N ∈ L(H) is increasing.
Proof. Let M := supkxk≤1 |hT x , xi|. Obviously M ≤ kT k because for kxk ≤ 1
Theorem 4.49. Let H be a Hilbert space. Every monotonic bounded sequence
|hT x , xi| ≤ kT kkxk2 ≤ kT k. of selfadjoint linear operators on H converges strongly.

To show the reverse inequality fix x, y ∈ H. Observe that Proof. Let (Tn )n∈N be a bounded monotonic sequence of selfadjoint operators.
Without restriction we assume that it is increasing. Let
hT (x + y) , x + yi − hT (x − y) , x − yi = 2hT x , yi + 2hT y , xi
= 2hT x , yi + 2hy , T xi = 4 RehT x , yi. snm : H × H → K, snm (x, y) = h(Tn − Tm )x , yi
84 4.5. Projections in Hilbert spaces Chapter 4. Hilbert spaces 85

is a positive semidefinite sesquilinear form on H if n ≥ m. Let M be a bound (v) =⇒ (i) Let x ∈ ker P , y ∈ rg P . Since for all λ ∈ R
of (Tn )n∈N . Note that then kTn − Tm k ≤ 2M . Then, using Cauchy-Schwarz
inequality, we find for n ≥ m and x ∈ H with kxk = 1 0 ≤ hP (x + λy) , x + λyi = hλy , x + λyi = λ2 kyk2 + λhy , xi,

k(Tn − Tm )xk2 = h(Tn − Tm )x , (Tn − Tm )xi = snm (x, (Tn − Tm )x) it follows that hx , yi = 0.
1 1
≤ snm (x, x) 2 snm ((Tn − Tm )x, (Tn − Tm )x) 2 Lemma 4.51. Let H Hilbert space H. A linear operator P : H → H is an
1
= h(Tn − Tm )x , xi 2 h(Tn − Tm )x , (Tn − Tm )2 xi 2
1
orthogonal projection if and only if P 2 = P and hx , P yi = hy , P xi for all
1 1 x, y ∈ H.
≤ h(Tn − Tm )x , xi 2 kTn − Tm k 2 kTn − Tm k
Proof. Assume that P is an orthogonal projection. Then P 2 = P and by Propo-
3 1
≤ (2M ) h(Tn − Tm )x , xi .
2 2

sition 4.50 P is selfadjoint.


By assumption (hTn x , xi)n∈N is a monotonically increasing bounded sequence If P 2 = P and hx , P yi = hy , P xi for all x, y ∈ H, then P is a projection. By
in R, hence convergent. It follows that (Tn x)n∈N is a Cauchy sequence, hence T the theorem of Hellinger-Toeplitz (Theorem 4.43) P is selfadjoint, hence P is
converges strongly to some T ∈ L(H) (Proposition 3.13). That T is selfadjoint an orthogonal projection by Proposition 4.50.
follows from
Lemma 4.52. Let H be a Hilbert space, U1 , U2 ⊆ H closed subspaces and P1 ,
hT x , yi = lim hTn x , yi = lim hx , Tn yi = hx , T yi, x, y ∈ H.
n→∞ n→∞ P2 the corresponding orthogonal projections. Then the following is equivalent:
(i) P1 P2 = P2 P1 = 0.
4.5 Projections in Hilbert spaces
(ii) U1 ⊥ U2 .
Proposition 4.50. Let H be a Hilbert space, P ∈ L(H) a projection. If P 6= 0 (iii) P1 + P2 is an orthogonal projection.
then the following is equivalent.
If one of the equivalent conditions above hold, then rg(P1 + P2 ) = U1 ⊕ U2 .
(i) P is an orthogonal projection.
(ii) kP k = 1. Proof. (i) =⇒ (ii) By assumption, U2 = rg P2 ⊆ ker P1 = (rg P1 )⊥ = U1⊥ ,
hence U1 ⊥ U2 .
(iii) P is selfadjoint.
(ii) =⇒ (i) By assumption, rg P2 = U2 ⊆ U1⊥ = ker P1 , hence P1 P2 = 0. Since
(iv) P is normal.
(ii) is symmetric in U1 and U2 , it follows also that P2 P1 = 0.
(v) hP x , xi ≥ 0, x ∈ H. (i),(ii) =⇒ (iii) Observe that P1 P2 = P2 P1 = 0, so P1 + P2 is a projection
because
Proof. (i) =⇒ (ii) follows from Theorem 4.16.
(ii) =⇒ (i) Let x ∈ ker P and y ∈ rg(P ). Then for all λ ∈ K (P1 + P2 )2 = P12 + P1 P2 + P2 P1 + P22 = P1 + P2 .

kλyk2 = kP (x + λy)k2 ≤ kx + λyk2 = kxk2 + |λ|2 kyk2 + 2 Re(λhx , yi). Since the sum of two selfadjoint operators is selfadjoint, P1 + P2 is selfadjoint,
hence, by Proposition 4.50 an orthogonal projection.
2 2
In particular, 0 ≤ kxk + 2λ Rehx , yi for all λ ∈ R, and 0 ≤ kxk + 2iλ Imhx , yi (iii) =⇒ (i) Since P1 + P2 is an orthogonal projection, it follows that
for all λ ∈ iR, hence Rehx , yi = Imhx , yi = 0.
P1 P2 + P2 P1 = (P1 + P2 )2 − (P1 + P2 ) = 0.
(i) =⇒ (iii) Observe that hP x , yi = hx , P yi for all x, y ∈ H because
In particular 0 = (P1 P2 +P2 P1 )P2 x = (id +P2 )P1 P2 x. Note that for y ∈ H \{0}
hP x , yi = hP x , y − P y + P yi = hP x , P yi,
the vectors (id −P2 )y and P2 y are linearly independent, hence (id +P2 )y =
hx , P yi = hx − P x + P x , P yi = hP x , P yi. (id −P2 )y + 2P2 y is zero if and only if (id −P2 )y = 0 and P2 y = 0, hence y = 0.
Therefore rg P1 P2 ⊆ ker(id +P2 ) = {0}.
(iii) =⇒ (iv) is clear.
(iv) =⇒ (i) By Lemma 4.47, ker P = ker P ∗ = (rg P )⊥ . Lemma 4.53. Let H be a Hilbert space and P1 and P2 orthogonal projections
(i) =⇒ (v) For all x ∈ H: hP x , xi = hP x , x − P x + P xi = hP x , P xi ≥ 0. on subspaces U1 and U2 .
86 4.5. Projections in Hilbert spaces Chapter 4. Hilbert spaces 87

(i) P1 P2 is an orthogonal projection if and only if P1 P2 = P2 P1 . In this case, Proof. By Theorem 4.49 we already know that s- lim Pn =: P exists and is a
P1 P2 is an projection on U1 ∩ U2 . selfadjoint operator. It remains to be shown that P is a projection, that is, that
P 2 = P . For x ∈ H and n ∈ N
(ii) P1 − P2 is an orthogonal projection if and only if P1 P2 = P2 P1 = P2 .
P 2 x = (P − Pn + Pn )(P − Pn + Pn )x = (P − Pn )P x + Pn (P − Pn )x + Pn2 x.
Proof. (i) If P1 P2 is an orthonormal projection, then, by Proposition 4.50,
P1 P2 is selfadjoint, that is P1 P2 = (P1 P2 )∗ = P2∗ P1∗ = P2 P1 . On the other Note that (P − Pn )P x → 0, n → ∞, and also Pn (P − Pn )x because kPn k = 1,
hand, if P1 and P2 commute, then it is easy to verify that (P1 P2 )2 = P1 P2 n ∈ N. Since Pn2 x = Pn x → P x, it follows that P 2 = P .
and (P1 P2 )∗ = P1 P2 , hence P1 P2 is an orthogonal projection. In this case,
rg(P1 P2 ) = rg(P2 P1 ), so rg(P1 P2 ) ⊆ U1 ∩ U2 . On the other hand, P1 P2 x = x
for every x ∈ U1 ∩ U2 , so also rg(P1 P2 ) ⊇ U1 ∩ U2 holds.
4.6 The adjoint of an unbounded operator
(ii) Using Lemma 4.52 we obtain In sections 2.4 and section 4.4 we have defined the adjoint of bounded linear
operators between Banach or Hilbert spaces. Now we define the adjoint of an un-
P1 − P2 orthonormal projection ⇐⇒ 1 − (P1 − P2 ) orthonormal projection bounded linear operator. Recall that T (X → Y ) denotes a possibly unbounded
⇐⇒ (1 − P1 ) + P2 orthonormal projection linear operators defined on a subspace D(T ) ⊆ X.
⇐⇒ P2 (1 − P1 ) = (1 − P1 )P2 = 0
Definition 4.56. Let X, Y be Banach spaces and D(T ) ⊆ X a dense subspace.
⇐⇒ P2 P1 = P1 P2 = P2 . For a linear map T : X ⊇ D(T ) → Y we define
Lemma 4.54. Let H be a Hilbert space and P1 , P2 orthogonal projections on D(T ′ ) := {ϕ ∈ Y ′ : x 7→ ϕ(T x) is a bounded linear functional on D(T )},
H0 , H1 ⊆ H. Then the following is equivalent.
Since D(T ) is dense in X, the map D(T ) → K, x 7→ ϕ(T x) has a unique
(i) H0 ⊆ H1 , continuous extension T ′ ϕ ∈ X ′ for ϕ ∈ D(T ′ ). Hence the Banach space adjoint
(ii) kP0 xk ≤ kP1 xk, x ∈ H. T′

(iii) hP0 x , xi ≤ hP1 x , xi, x ∈ H. T ′ : Y ′ ⊇ D(T ′ ) → X ′ , (T ′ ϕ)(x) = ϕ(T x), x ∈ D(T ), ϕ ∈ D(T ′ ).
(iv) P0 P1 = P0 . is well-defined.
Proof. (ii) ⇐⇒ (iii) Let x ∈ H and P an orthogonal projection. Then hP x , xi = Theorem 4.57. Let X, Y be Banach spaces, D(T ) ⊆ X a dense subspace and
hP 2 x , xi = hP x , P xi = kP xk2 . T : X ⊇ D(T ) → Y be a linear operator. Then T ′ is closed.
(i) ⇐⇒ (iv)
Proof. Let G(T ′ ) = {(y ′ , T ′ y ′ ) : ϕ ∈ D(T ′ )} ⊆ Y ′ × X ′ be the graph of T ′ .
P0 P1 = P0 ⇐⇒ P0 (id −P1 ) = 0 ⇐⇒ rg(id −P1 ) ⊆ ker P0 Note that (y ′ , x′ ) ∈ G(T ′ ) if and only if x′ x = y ′ (T x) for all x ∈ D(T ). Now let
⇐⇒ (rg P1 )⊥ ⊆ (rg P0 )⊥ ⇐⇒ H1⊥ ⊆ H0⊥ ((yn′ , x′n ))n∈N ⊆ G(T ′ ) a convergent sequence with lim (yn′ , x′n ) = (y0′ , x′0 ). For
n→∞
⇐⇒ H0 ⊆ H1 . all x ∈ D(T ) it follows that

(iv) =⇒ (ii) For all x ∈ H: kP0 xk = kP0 P1 xk ≤ kP0 kkP1 xk ≤ kP1 xk. x′0 x = lim x′n x = lim yn′ (T x) = lim y0′ (T x),
n→∞ n→∞ n→∞

(iii) =⇒ (i) Let x ∈ H1⊥ = ker P1 . Then 0 = hP1 x , xi ≥ hP0 x , xi ≥ 0, hence thus (y0′ , x′0 ) ∈ G(T ′ ) which implies that T ′ is closed.
hP0 x , xi = 0. It follows that P0 |H1⊥ = 0 (Corollary 4.46), hence H1⊥ ⊆ ker P0 =
H0⊥ . Definition 4.58. Let X, Y be Banach spaces. For linear operators S, T from
X to Y we write S ⊆ T if T is an extension of S, that is, if D(S) ⊆ D(T ) and
Lemma 4.55. Let H be a Hilbert space and (Pn )n∈N a sequence of orthogonal T |D(S) = S.
projections with hPm x , xi ≤ hPn x , xi for all x ∈ X and m < n. Then (Pn )n∈N
converges strongly to an orthogonal projection. Theorem 4.59. Let X, Y, Z be Banach spaces.
88 4.6. The adjoint of an unbounded operator Chapter 4. Hilbert spaces 89

(i) Let (S, D(S)) and (T, D(T )) be densely defined linear operators X → Y . Remark 4.62. The sets A◦ and ◦ B are closed subspaces and ◦ (A◦ ) = A. If X
If S ⊆ T then T ′ ⊆ S ′ . is reflexive, then also (◦ B)◦ = B.
(ii) Assume S(X → Y ) and T (Y → Z) are densely defined such that also T S
is densely defined. Then S ′ T ′ ⊆ (T S)′ . Proof. Obviously, A◦ and ◦ B are subspaces. Let (x′n )n∈N ⊆ A◦ be a convergent
(iii) Assume S(X → Y ) and T (X → Y ) are densely defined such that also sequence. Then x′0 := lim x′n ∈ A◦ because x′0 x = lim x′n x = 0 for all x ∈ A.
n→∞ n→∞
T + S is densely defined. Then (S ′ + T ′ ) ⊆ (S + T )′ . Let (xn )n∈N ⊆ ◦ B be a convergent sequence. Then x0 := lim xn ∈ ◦ B because
n→∞
ϕx0 = lim ϕxn = 0 for all ϕ ∈ B.
n→∞
Proof. (i) is clear from the definition of the adjoint operator.
(ii) Let z ′ ∈ D(S ′ T ′ ). Then T ′ z ′ ∈ D(S ′ ) and the map Now we show that ◦ (A◦ ) = A. Since obviously A ⊆ ◦ (A◦ ), also A ⊆ ◦ (A◦ ).
Assume that there exists an a ∈ ◦ (A◦ ) \ A. By a corollary to the Hahn-Banach
D(S) → K, x 7→ (T ′ z ′ )(Sx) theorem (Corollary 2.19) there exists a ϕ ∈ X ′ such that ϕ|A = 0 and ϕ(a) 6= 0.
Therefore ϕ ∈ A◦ , so by definition of ◦ (A◦ ), also ϕ(a) = 0.
is continuous. Then also its restriction
(◦ B)◦ = B follows if we identify X with X ′′ using the canonical map JX .
D(T S) → K, x 7→ (T ′ z ′ )(Sx) = z ′ (T Sx)
is continuous. Note that by assumption D(T S) is dense in X, hence z ′ ∈ Lemma 4.63. Let X, Y be Banach space, Y 6= {0} and T (X → Y ) a densely
D((T S)′ ) and (T S)′ z ′ = S ′ T ′ z ′ . defined closed linear operator and y0 ∈ Y \ {0}. Then there exists a ϕ ∈ D(T ′ )
such that ϕ(y0 ) 6= 0, in particular, D(T ′ ) 6= {0}.
(iii) Let y ′ ∈ D(T ′ + S ′ ) = D(T ′ ) ∩ D(S ′ ). Then the map
D(T + S) → K, x 7→ y ′ (T x) + y ′ (Sx) = y ′ ((T + S)x) Proof. By assumption, the graph G(T ) of T is closed and (0, y0 ) 6= G(T ). Hence,
by a corollary to the Hahn-Banach theorem (Corollary 2.19) there exists ψ ∈
is continuous. Since by assumption D(T + S) is dense in X, y ′ ∈ D((T + S)′ )
(X × Y )′ such that ψ|G(T ) = 0 and ψ((0, y0 )) 6= 0. Let ϕ : Y → K, ϕ(y) =
and (T + S)′ y ′ = (T ′ + S ′ )y ′ .
ψ((0, y)). Obviously ϕ ∈ Y ′ and ϕ(y0 ) 6= 0. Moreover, ϕ ∈ D(T ′ ) because for
If S and T are bounded, then “=” holds in (ii) and (iii) (Theorem 2.26). Note all x ∈ D(T )
that for unbounded linear operators T ′ +S ′ = (T +S)′ is not necessarily true. For
example, if T (X → Y ) is a densely defined unbounded linear operator such that ϕ(T x) = ψ((0, T x)) = ψ((x, T x) − (x, 0)) = ψ((x, T x)) − ψ((x, 0))
also T ′ is densely defined with D(T ′ ) 6= Y ′ . Then D(T ′ − T ′ ) 6= Y ′ = D(T − T )′ . = −ψ((x, 0)).

Corollary 4.60. Let X be a Banach space, T a densely defined linear operator Theorem 4.64. Let X and Y be Banach spaces. For a densely defined closed
in X with bounded inverse T −1 ∈ L(X). Then T ′ is invertible and linear operator T (X → Y ) the following holds:

(T ′ )−1 = (T −1 )′ . (i) rg(T )◦ = rg(T ) = ker T ′ .
◦ ′
(ii) rg T = (ker T ).
Proof. By Theorem 4.59 (ii) it follows that (T −1 )′ T ′ ⊆ (T T −1)′ = id′X = idX ,
hence (T −1 )′ T ′ = idD(T ′ ) . (iii) rg T = Y ⇐⇒ T ′ is injective.
Again by Theorem 4.59 (ii) we find T ′ (T −1 )′ ⊆ (T −1 T )′ = id′D(T ) = idX , so it (iv) ◦ (rg T ′ ) ∩ D(T ) = ker T.
suffices to show D(T ′ (T −1 )′ ) = D(T ′ ). Let ϕ ∈ D(T ′ ) and η = (T −1 )′ ϕ. For (v) rg T ′ ⊆ (ker T )◦ .

every x ∈ D(T ) it follows that η(T x) = ((T −1 )′ ϕ)(T x) = ϕ(T −1 T x) = ϕ(x),
which implies η ∈ D(T ′ ), hence D(T ′ (T −1 )′ ) = D(T ′ ). Proof. (i) The first equality is clear. The second equality follows from
More general is Theorem 4.65 due to Phillips. ϕ ∈ rg(T )◦ ⇐⇒ ∀ y ∈ rg(T ) ϕ(y) = 0
Definition 4.61. Let X be a Banach space. For subspaces A ⊆ X and B ⊆ X ′ ⇐⇒ ∀ x ∈ D(T ) ϕ(T x) = 0
we define the annihilators ⇐⇒ ϕ ∈ D(T ′ ), T ′ ϕ = 0
A◦ := {ϕ ∈ X ′ : ϕ(x) = 0, x ∈ A} ⊆ X ′ , ⇐⇒ ϕ ∈ ker(T ′ ).

B := {x ∈ X : ϕ(x) = 0, ϕ ∈ B} ⊆ X. (ii) rg T = ◦ ((rg T )◦ ) = ◦ (ker T ′ ) by (i) and Remark 4.62.
90 4.6. The adjoint of an unbounded operator Chapter 4. Hilbert spaces 91

(iii) By (ii), rg T = Y if and only if ◦ (ker T ′ ) = Y . This is the case if and only (i) ⇐⇒ (v) follows from theorem 4.64 (ii).
if ϕ(y) = 0 for all ϕ ∈ ker T ′ and y ∈ Y , that is, if and only if ker T ′ = {0}. (ii) ⇐⇒ (vi) follows from theorem 4.64 (ii)
(iv) Let x ∈ ker(T ) and x′ ∈ rg T ′ . Choose y ′ ∈ D(T ′ ) with T ′ y ′ = x′ . Then
rg(T ′ ) = ◦ (ker T ′′ ) = (ker T )◦ .
x′ x = (T ′ y ′ )x = y ′ (T x) = y ′ (0) = 0, hence x ∈ ◦ (rg T ′ ).
Now let x ∈ ◦ (rg T ′ ) ∩ D(T ). Then y ′ (T x) = (T ′ y ′ )x = 0 for all y ′ ∈ Y ′ . Since (iii) ⇐⇒ (iv) Recall that T is open if and only if there exists an , r > 0 such
T is closed, it follows by Lemma 4.63 that T x = 0, hence x ∈ ker T . that the image of the open ball in X with centre 0 and radius r contains the open
′ ′ ′ ′ ′ ′ ′
(v) Let x ∈ rg(T ) and x ∈ ker T . Choose y ∈ D(T ) such that T y = x . unit ball in Y . That is, there exists a r > 0 such that T (BX (0, r)) ⊇ BY (0, 1).
Then x′ x = (T ′ y ′ )x = y ′ (T x) = y ′ (0) = 0. It follows that rg(T ′ ) ⊆ (ker T )◦ , Assume that T is open and let r as above.
and since (ker T )◦ is closed, the statement is proved. To show that T ′ is open, we have to show that for every x′0 ∈ rg(T ′ ) with
kx′0 k < 1, there exists a y0′ ∈ D(T ′ ) with T ′ y0′ = x′0 and ky0′ k < r. Define
Theorem 4.65 (Phillips). Let X, Y be a Banach spaces, T (X → Y ) a densely a linear functional ϕ on rg(T ) as follows: for y ∈ rg T with kyk < 1 choose
defined injective linear operator with rg(T ) = Y . Then x ∈ D(T ) such that kxk < r and T x = y. Set ϕ(y) = x′0 x and extend ϕ linearly
to rg T . Note that |ϕ(y)| = |x′0 x| ≤ kx′0 kkxk ≤ rkyk, ϕ is bounded, so by
(T ′ )−1 = (T −1 )′ (4.2) the theorem of Hahn-Banach it can be extended to a functional y0′ ∈ Y ′ with
ky0′ k ≤ r. Note that
and T −1 is bounded if and only if T is closed and (T ′ )−1 is bounded on X ′ .
(T −1 denotes the inverse of T : D(T ) → rg(T ), similar for (T ′ )−1 .) D(T ) → K, 7→ y0′ (T x) = ϕ(T x) = x′0 x

Proof. is continuous, so y0′ ∈ D(T ).


(iv) ⇐⇒ (iii) Follows analogously if we note that T ′′ = T by the reflexivity of
Theorem 4.66 (Closed range theorem). Let X, Y be reflexive Banach spaces X and Y .
and T : X ⊇ D(T ) → Y a closed densely defined linear operator. The following
is equivalent: Definition 4.67. Let H1 , H2 be Hilbert spaces and D(T ) ⊆ H1 a dense sub-
space. For a linear map T : H1 ⊇ D(T ) → H2 its Hilbert space adjoint T ∗ is
(i) rg(T ) is closed.
defined by
(ii) rg(T ′ ) is closed.
D(T ∗ ) := {y ∈ H2 : x 7→ hT x , yi is a bounded on D(T )},
(iii) T : X ⊇ D(T ) → rg(T ) is open.
T ∗ : H2 ⊇ D(T ∗ ) → H1 , T ∗y = y∗ ,
(iv) T ′ : Y ′ ⊇ D(T ′ ) → rg(T ′ ) is open.
(v) rg(T ) = ◦ (ker T ′ ). where y ∗ ∈ H1 such that hT x , yi = hx , y ∗ i for all x ∈ D(T ).
′ ◦ Note that for y ∈ D(T ∗ ) the map x 7→ hT x , yi is continuous and densely defined
(vi) rg(T ) = (ker T ) .
and can therefore be extended uniquely to an element ϕy ∈ H1′ . By the Riesz
representation theorem (Theorem 4.20) there exists exactly one y ∗ ∈ H1 as
Proof. (i) ⇐⇒ (iii) Since T is closed, (D(T ), k · kT ) is a Banach space and
desired.
Te : (D(T ), k · kT ) → rg T, Tex = T x
Definition 4.68. Let H1 , H2 be Hilbert spaces and D(T ) ⊆ H1 , D(S) ⊆ H2
is continuous (Lemma 3.32). Observe that also i : (D(T ), k · kT ) → X, x 7→ x is subspaces. The linear maps T : H1 ⊇ D(T ) → H2 and S : H2 ⊇ D(S) → H1
continuous and that T = Te ◦ i−1 : X ⊇ D(T ) → Y . Note that rg T is a Banach are called formally adjoint if
space. hT x , yiH2 = hx , SyiH1 , x ∈ D(T ), y ∈ D(S).
If rg T is closed, then Te : (D, k · kT ) → rg T is open by the open mapping
theorem (Theorem 3.22), then also T = Te ◦ i−1 : X ⊇ D(T ) → rg T is open as Note that the formal adjoint of a non-densely defined linear operator is not
composition of open maps. If T : D(T ) → rg T is open, then it is surjective, unique; in particular, the operator trivial operator with D = {0} is formally
hence rg T is closed. adjoint to every linear operator.
Note that T ′ is closed (Theorem 4.57), hence (ii) ⇐⇒ (iv) is proved analogously. If T is densely defined, then its adjoint T ∗ is its maximal formally adjoint
operator.
92 4.6. The adjoint of an unbounded operator Chapter 4. Hilbert spaces 93

Lemma 4.69. Let H1 and H2 be Hilbert spaces and define (ii) Assume S(H1 → H2 ) and T (H2 → H3 ) are densely defined with T S = H1 .
Then S ∗ T ∗ ⊆ (T S)∗ .
U : H1 × H2 → H2 × H1 , (x, y) 7→ (y, −x).
(iii) Assume S(H1 → H2 ) and T (H1 → H2 ) are densely defined with T + S =
If T (H1 → H2 ) is a densely defined linear operator, then H1 . Then (T ∗ + S ∗ ) ⊆ (S + T )∗ .
G(T ∗ ) = U (G(T )⊥ ) = [U (G(T ))]⊥ . (4.3) If S and T are bounded, then “=” holds in (ii) and (iii).
⊥ ⊥
Proof. Observe that U is unitary, hence U (G(T ) ) = [U (G(T ))] . The first Proof. As is in the Banach space case.
equality in (4.3) follows from
Corollary 4.72. Let H be a Hilbert space, T a densely defined linear operator
(y0 , x0 ) ∈ G(T ∗ ) ⇐⇒ hT x , y0 iY = hx , x0 iX , x ∈ D(T ) in H with bounded inverse T −1 ∈ L(H). Then T ∗ is invertible and
⇐⇒ hT x , y0 i − hx , x0 i = 0, x ∈ D(T )
⇐⇒ h(T x, −x) , (y0 , x0 )iH2 ×H1 = 0, x ∈ D(T ) (T ∗ )−1 = (T −1 )∗ =: T −∗ .
⇐⇒ hU (x, T x) , (y0 , x0 )iH2 ×H1 = 0, x ∈ D(T )
Proof. By Theorem 4.71 (ii) it follows that (T −1 )∗ T ∗ ⊆ (T T −1 )∗ = idH ∗ = idH ,
⇐⇒ (y0 , x0 ) ∈ [U (G(T ))]⊥ . hence (T −1 )∗ T ∗ = idD(T ∗ ) .
Again by Theorem 4.71 (ii) we find T ∗ (T −1 )∗ ⊆ (T −1 T )∗ = id∗D(T ) = idH , so it
Theorem 4.70. Let H1 and H2 be Hilbert spaces. For a densely defined linear suffices to show D(T ∗ (T −1 )∗ ) = D(T ∗ ). Let y ∈ D(T ∗ ) and z = (T −1 )∗ y. For
operator T (X → Y ) the following holds: every x ∈ D(T ) it follows that hT x , zi = hT x , (T −1 )∗ yi = hT −1 T x , yi = hx , yi,
(i) T ∗ is closed. so z ∈ D(T ∗ ) which implies D(T ∗ (T −1 )∗ ) = D(T ∗ ).
(ii) If T is closable, then T ∗ is densely defined and T ∗∗ = T .
Theorem 4.73. Let H1 , H2 be Hilbert spaces, T (H1 → H2 ) a densely defined
closed linear operator. Then the following holds.
Proof. (i) follows immediately from (4.3).

(ii) Let y0 ∈ D(T ∗ )⊥ . Then hy0 , yi = 0 for all y ∈ D(T ). This implies (i) rg(T )⊥ = rg(T ) = ker T ∗ .
0 = h(0, y0 ) , (−z, y)iH1 ×H2 = h(0, y0 ) , U (y, z)iH1 ×H2 , (y, z) ∈ G(T ∗ ). (ii) rg(T ) = (ker T ∗ )⊥ .
Hence by Lemma 4.69, (iii) rg(T ∗ )⊥ = ker T .
−1 ∗ ⊥ ⊥⊥
(0, y0 ) ∈ [U (G(T ))] = G(T ) = G(T ) = G(T ). (iv) rg(T ∗ ) = (ker T )⊥ .

It follows that y0 = T 0 = 0, so D(T ∗ ) = Y . Let Proof. (i) Note that y ∈ rg(T )⊥ if and only if hT x , yi for all x ∈ D(T ). This
is equivalent to y ∈ D(T ∗ ) and T ∗ y = 0.
V : H 2 × H1 → H 1 × H2 , V (y, x) = (x, −y).
⊥⊥
(ii) By (i) rg(T ) = rg(T ) = (ker T ∗ )⊥ .
Obviously V U = − idH1 ×H2 and application of Lemma 4.69 to T ∗ yields
(iii) By Theorem 4.70 T is closed and densely defined and T ∗∗ = T . Appli-

G(T ∗∗ ) = [V (G(T ∗ ))]⊥ = [V U (G(T )⊥ )]⊥ = [−(G(T )⊥ )]⊥ = G(T )⊥⊥ = G(T ) cation of (i) to T ∗ shows rg(T ∗ )⊥ = ker T .
= G(T ). (iv) Application of (ii) to T ∗ shows rg(T ∗ ) = (ker T )⊥ .

hence T ∗∗ = T . Example 4.74. Let H = L2 [0, 1]. Let

Theorem 4.71. Let H1 , H2 , H3 be Hilbert spaces. D(T1 ) := W21 (0, 1) = {x ∈ L2 [0, 1] : x absolutely continuous, x′ ∈ L2 [0, 1]},
D(T2 ) := D(T1 ) ∩ {x ∈ L2 [0, 1] : x(0) = x(1)}
(i) Let T (H1 → H2 ) and S(H1 → H2 ) be densely defined linear operators. If
S ⊆ T then T ∗ ⊆ S ∗ . D(T3 ) := D(T1 ) ∩ {x ∈ L2 [0, 1] : x(0) = x(1) = 0}.
94 4.6. The adjoint of an unbounded operator Chapter 4. Hilbert spaces 95

For k = 1, 2, 3 let (ii) T is called selfadjoint if T = T ∗ .


′ (iii) T is called essentially selfadjoint if T = T ∗ .
Tk : H ⊇ D(Tk ) → H, Tk x = ix .

Obviously, the Tk are well-defined and D(Tk ) is dense in H (Theorem A.27). The operator T2 in the example above is selfadjoint, the operator T3 is symmet-
We will show: T1∗ = T3 , T3∗ = T1 , T2∗ = T2 , in particular all Tk are closed. ric.

Proof. Let x, y ∈ D(T1 ). Then, using integration by parts,


Z 1 1 Z 1

hT1 x , yi = ix′ (t)y(t) dt = ix(t)y(t) − ix(t)y ′ (t) dt
0 0 0
= ix(1)y(1) − ix(0)y(0) + hx , T1 yi.

In particular we obtain

hT x , yi = hx , T yi, x ∈ D(T1 ), y ∈ D(T3 ),


hT x , yi = hx , T yi, x, y ∈ D(T2 ).

This shows that

D(T3 ) ⊆ D(T1∗ ), D(T2 ) ⊆ D(T2∗ ) and D(T1 ) ⊆ D(T3∗ )

and T1∗ |D(T3 ) = T3 , T3∗ |D(T3 ) = T1 and T2∗ |D(T3 ) = T2 .


To prove the inclusion D(T1∗ ) ⊆ D(T3 ) let g ∈ D(T1∗ ) and ϕ = T1∗ g. Define
Rt
Φ(t) = 0 ϕ(s) ds. Then Φ is absolutely continuous and Φ′ = ϕ. For x ∈ D(T1 )
Z 1 Z 1
ix′ (t)g(t) dt = hT1 x , gi = hx , ϕi = ix(t)ϕ(t) dt
0 0
1 Z 1

= x(t)Φ(t) − ix′ (t)Φ(t) dt
0 0
Z 1
= x(1)Φ(1) − ix′ (t)Φ(t) dt.
0

Note that Φ(1) = 0 as can be seen if x is chosen to be a constant function.


Hence
Z 1
ix′ (t)(g(t)iΦ(x)) dt = 0, x ∈ D(T1 ),
0

implying that g + iΦ ∈ rg(T1 )⊥ = {0}. It follows that g is absolutely continuous


and g(0) = iϕ(0) = 0, g(1) = iϕ(1) = 0, so g ∈ D(T3 ).
Analogously, T2∗ = T2 and T3∗ = T1 can be shown.

Definition 4.75. Let H be a Hilbert spaces, D(T ) ⊆ H a dense subspace and


T : H ⊇ D(T ) → H a linear map.
(i) T is called symmetric if T ⊆ T ∗ .
Chapter 5. Spectrum of linear operators 97

Chapter 5

Spectrum of linear
operators

If not stated explicitely otherwise, all Hilbert and Banach spaces in this chapter
are assumed to be complex vector spaces.

5.1 The spectrum of a linear operator


Definition 5.1. Let X be a Banach space and T (X → X) a densely defined
linear operator.

ρ(T ) := {λ ∈ C : λ id −T is bijective} resolvent set of T,


σ(T ) := C \ ρ(T ) spectrum of T.

The spectrum of T is further divided in point spectrum σp (T ), continuous spec-


trum σc (T ) and residual spectrum σr (T ):

σp (T ) := {λ ∈ C : λ id −T is not injective},
σc (T ) := {λ ∈ C : λ id −T is injective, rg(T − λ id) 6= X, rg(T − λ id) = X},
σr (T ) := {λ ∈ C : λ id −T is injective, rg(T − λ id) 6= X}.

It follows immediately from the definition that

σ(T ) = σp (T ) ∪˙ σc (T ) ∪˙ σr (T ).

In the following, we often write λ − T instead of λ id −T .

Definition 5.2. (i) Elements λ ∈ σp (T ) are called eigenvalues of T .

(ii) For λ ∈ σp (T ) we define the geometric eigenspace of T in λ, Nλ (T ), and


98 5.2. The resolvent Chapter 5. Spectrum of linear operators 99

the algebraic eigenspace of T in λ, Aλ (T ), by Lemma 5.7. Let X be a Banach space and T (X → X) a closed linear operator.

Nλ (T ) := ker(T − λ), 1
(i) kR(λ0 , T )k ≥ for all λ0 ∈ ρ(T ).
dist(λ0 , σ(T ))
Aλ (T ) := {x ∈ X : (T − λ)n x = 0 for some n ∈ N}.
(ii) For λ0 ∈ ρ(T ) and λ ∈ C with |λ − λ0 | < kR(λ0 , T )k−1
(iii) For λ ∈ ρ(T ) the resolvent of T in λ is (λ id −T )−1 := R(λ, T ). The map ∞
X
R(λ, T ) = (λ0 − λ)n (R(λ0 , T ))n+1 .
ρ(T ) → L(X), λ 7→ R(λ, T ) n=0

is the resolvent map. Note that (ii) shows that locally around a λ0 ∈ ρ(T ) the resolvent has a power
−1
series expansion with coefficients depending only on λ0 and T .
Remark 5.3. If T is closed, then (T − λ) is closed if it exists. Therefore, by
the closed graph theorem, Proof of Lemma 5.7. Recall that for a bounded linear operator S ∈ L(X) with
kSk < 1 the operator (id −S)−1 ∈ L(X) and it is given explicitly by the Neu-
−1
ρ(T ) = {λ ∈ C : T − λ is injective and (T − λ) ∈ L(X)}. mann series (Theorem 2.10)

X
Remark 5.4. Often the resolvent set of a linear operator is defined slightly
(id −S)−1 = Sn.
differently: Let T (X → X) is a densely defined linear operator. Then λ ∈ ρ(T )
n=0
if and only if λ − T is bijective and (λ − T )−1 ∈ L(X). With this definition
it follows that ρ(T ) = ∅ for every non-closed T (X → X) because one of the Let λ0 ∈ ρ(T ). For λ ∈ C we find
following cases holds:  
λ − T = λ0 − T − (λ0 − λ) = id −(λ0 − λ)(λ0 − T )−1 (λ0 − T ).
(i) λ − T is not bijective =⇒ λ ∈ / ρ(T );
If |λ0 − λ| < k(λ0 − T )−1 k−1 , then the term in brackets is invertible, hence so
(ii) λ − T is bijective, then (λ − T )−1 is defined everywhere and not closed, so
is λ − T and we obtain
it cannot be bounded, which implies λ ∈ / ρ(T ).
 −1
(λ − T )−1 = (λ0 − T )−1 id −(λ0 − λ)(λ0 − T )−1
Remark 5.5. If dim X < ∞, then σc (T ) = σr (T ) = ∅ and σp (T ) is the set of X ∞ 
all eigenvalues of T . = (λ0 − T )−1 (λ0 − λ)n (λ0 − T )−n
n=0
Theorem 5.6 (Spectral mapping theorem for polynomials). Let X be a ∞
X
Banach space, T ∈ L(X) and P ∈ C[X] a polynomial. Then = (λ0 − λ) (λ0 − T )−(n+1)
n

n=0
σ(P (T )) = P (σ(T )).
which proves (ii). If µ ∈ C with |µ| < k(T − λ0 )−1 k−1 , then λ0 + µ ∈ ρ(T ),
Proof. Let λ ∈ C. Then there exists a polynomial Q such that P (X) − P (λ) = hence dist(λ0 , σ(T )) ≥ k(T − λ0 )−1 k−1 , so also (i) is proved.
(X − λ)Q(X). In particular, P (T ) − P (λ) = (T − λ)Q(T ) = Q(T )(T − λ). As a corollary we obtain the following theorem.
Hence, if λ ∈ σ(T ), then (T − λ) is not bijective, so P (T ) − P (λ) is not bijective
which implies P (σ(T )) ⊆ σ(P (T )). Theorem 5.8. Let X be a Banach space and T (X → X) a closed linear oper-
Now assume µ ∈ σ(P (T )). There exist a, λ1 , . . . , λn ∈ C such that P (X) − µ = ator.
a(X − λ1 ) · · · (X − λn ). Since P (T ) − µ is not invertible, at least one of the
terms λj − T cannot be invertible, that is at least one λj must belong to the (i) σ(T ) is closed.
spectrum of T and µ = P (λj ) ∈ P (σ(T )). (ii) If T ∈ L(X), then σ(T ) is compact.

5.2 The resolvent Proof. (i) C \ σ(T ) = ρ(T ) is open by Lemma 5.7.
(ii) Let λ ∈ C with |λ| > kT k. Then λ − T = λ(id −λ−1 T ) is invertible since
In this section we will study the resolvent map ρ(T ) → L(X), λ 7→ R(λ, T ) = kλ−1 T k < 1 (Neumann series, Theorem 2.10), hence λ ∈ ρ(T ) It follows that
(λ − T )−1 . We will show that its domain is open and that it is analytic. {λ ∈ C : |λ| > kT k} ⊇ ρ(T ). Since σ is closed and bounded, it is compact.
100 5.2. The resolvent Chapter 5. Spectrum of linear operators 101

Next we prove the so-called resolvent identities. Lemma 5.11. Let X be a Banach space. A sequence (xn )n∈N ⊆ X is a Cauchy
sequence if and only if the sequence (ϕ(xn ))n∈N ⊆ X is uniformly Cauchy for
Theorem 5.9. Let X be a Banach space and T (X → X), S(X → X) a linear ϕ ∈ X ′ with kϕk ≤ 1 (that is, for every ε > 0 exists a N ∈ N such that
operators with D(S) = D(T ). |ϕ(xn ) − ϕ(xm )| < ε for all m, n ≥ N and all ϕ ∈ X ′ with kϕk ≤ 1).
(i) 1st resolvent identity:
Proof. Assume that (xn )n∈N ⊆ X is a Cauchy sequence and let ε > 0. Then
R(λ, T ) − R(µ, T ) = (µ − λ)R(λ, T )R(µ, T ), λ, µ ∈ ρ(T ). there exists a N ∈ N such that kxn − xm k < ε for m, n ≥ N . It follows that
kϕ(xn ) − ϕ(xm )k ≤ kϕkkxn − xm k < ε for all m, n ≥ N and all ϕ ∈ X ′ with
In particular, the resolvents commute. kϕk ≤ 1.
(ii) 2nd resolvent identity: Now let ε > 0 and assume that there exists an N ∈ N such that |ϕ(xn ) −
ϕ(xm )| < ε for all m, n ≥ N and all ϕ ∈ X ′ with kϕk ≤ 1. Recall that the map
R(λ, T ) − R(λ, S) = R(λ, T )(T − S)R(λ, S), λ ∈ ρ(T ) ∩ ρ(S). JX : X → X ′′ is an isometry. It follows for m, n ≥ N

Proof. (i) follows from a straightforward calculation: kxn − xm k = kJX xn − JX xm k = sup{|(JX xn − JX xm )ϕ| : ϕ ∈ X ′ , kϕk ≤ 1}
R(λ, T ) − R(µ, T ) = (λ − T ) −1
− (µ − T ) −1 = sup{|ϕ(xn ) − ϕ(xm )| : ϕ ∈ X ′ , kϕk ≤ 1} < ε.
 
= (λ − T )−1 µ − T − (λ − T ) (µ − T )−1
Recall the following fundamental theorem of complex analysis.
= (µ − λ)R(λ, T )R(µ, T ).

(ii) is shown similarly: Theorem 5.12 (Cauchy’s integral formula). Let Ω ∈ C open and let f :
Ω → C holomorphic. Let z0 ∈ Ω and r > 0 such that Kr (z0 ) := {z ∈ C :
−1 −1
R(λ, T ) − R(λ, S) = (λ − T ) − (λ − S) |z − z0 | ≤ r} ⊆ Ω. Then
 
= (λ − T )−1 λ − S − (λ − T ) (λ − S)−1 Z
1 f (z)
= R(λ, T )(T − S)R(λ, S), f (a) = dz, a ∈ Br (z0 ) (5.1)
2πi Γr (z0 ) z−a
Next we study properties of the resolvent map ρ(T ) → L(X), λ 7→ R(λ, T ). By
Lemma 5.7 we already now that its domain is open and that it is analytic, that where Γr (z0 ) is the positively oriented boundary of Kr (z0 ). More generally, for
is, locally it has a power series representation. n ∈ N0 ,
Z
Definition 5.10. Let Ω ∈ C be an open set, X a Banach space and f : Ω → X. n! f (z)
f (n) (a) = dz, a ∈ Br (z0 ). (5.2)
(i) f is called holomorphic in z0 ∈ Ω if and only if the limit 2πi Γr (z0 ) (z − a)n+1

f (z) − f (z0 ) Theorem 5.13 (Dunford). Let X be a Banach space and let Ω ∈ C open. A
lim
z→z0 z − z0 map f : Ω → X is holomorphic if and only if it is weakly holomorphic.
exists in the norm topology. f is called holomorphic if and only if it is
holomorphic in every z0 ∈ Ω. Proof. Clearly, holomorphy of f implies weak holomorphy. Now assume that f
is weakly holomorphic. Let z0 ∈ Ω. Choose r > 0 such that Kr (z0 ) = {z ∈ C :
(ii) f is called weakly holomorphic in z0 ∈ Ω if and only if the limit |z − z0 | ≤ r} ∈ Ω. and let Γr (z0 ) be the positively oriented boundary of Kr (z0 ).
For every ϕ ∈ X ′ Cauchy’s integral formula (5.1) yields
f (z) − f (z0 )
lim Z
z→z0 z − z0 1 ϕ(f (z))
ϕ(f (a)) = dz, a ∈ Br (z0 ).
exists in the weak topology. f is called weakly holomorphic if and only if 2πi Γr (z0 ) z−a
it is weakly holomorphic in every z0 ∈ Ω. Hence, for every ϕ ∈ X ′ the
map Ω → C, z 7→ ϕ(f (z)) is holomorphic in the usual sense. For a ∈ Br (z0 ) and 0 < |h| < r − |z0 − a| it follows that a + h ∈ Kr (z0 ), hence
102 5.2. The resolvent Chapter 5. Spectrum of linear operators 103

with Cauchy’s integral formula we obtain Since z 7→ ϕ(T (z)x) is holomorphic in a neighbourhood of Γr (z0 ), it is continu-
ous, so there exists Cx,ϕ such that
1 
ϕ(f (a + h)) − ϕ(f (a)) − (ϕ ◦ f )′ (a) |ϕ(T (z)x)| < Cx,ϕ , z ∈ Γr (z0 ).
h Z
1 1h 1 1 h i
= − − ϕ(f (z)) dz By a corollary to the theorem of Banach-Steinhaus (Corollary 3.8), there exists
2πi Γr (z0 ) h z − a − h z − a (z − a)2 Cx > 0 such that
Z h i
1 1 1
= − ϕ(f (z)) dz kT (z)xk < Cx , z ∈ Γr (z0 ),
2πi Γr (z0 ) (z − a)(z − a − h) (z − a)2
Z
h ϕ(f (z)) and by the theorem of Banach-Steinhaus (Theorem 3.7), there exists C > 0
= dz.
2πi Γr (z0 ) (z − a)2 (z − a − h) such that

Since z 7→ ϕ(f (z)) is holomorphic in a neighbourhood of Γr (z0 ), it is in partic- kT (z)k < C, z ∈ Γr (z0 ).
ular continuous. Hence there exists Cϕ such that
This implies that
|ϕ(f (z))| < Cϕ , z ∈ Γr (z0 ). 1   1 
lim ϕ(T (a + h)x − T (a)x) = ϕ lim (T (a + h)x − T (a)x)
h→0 h h→0 h
By a corollary to the theorem of Banach-Steinhaus (Corollary 3.8), there exists
C > 0 such that exists, uniformly for ϕ ∈ X ′ , kϕk ≤ 1. Therefore, by Lemma 5.11,

kf (z)k < C, z ∈ Γr (z0 ). 1


lim (T (a + h)x − T (a)x)
h→0 h
Hence we obtain
exists and convergence is uniform for x ∈ X with kxk = 1. Analogously as in
1  d
the proof of Lemma 5.11 it follows the existence of
ϕ(f (a + h)) − ϕ(f (a)) − (ϕ ◦ f )(a) ≤ hkϕkC ′ .
h dz
1
lim (T (a + h) − T (a)).
This implies that h→0 h
1  1 
lim ϕ f (a + h) − f (a) = lim ϕ(f (a + h)) − ϕ(f (a)) = (ϕ ◦ f )′ (a), Theorem 5.15. Let X be a Banach space, T (X → X) a densely defined closed
h→0 h h→0 h linear operator. Then the resolvent map
uniformly for ϕ ∈ X ′ , kϕk ≤ 1. Therefore, by Lemma 5.11, lim h1 f (a + h) − ρ(T ) → L(X), λ 7→ R(λ, T ) = (λ − T )−1
 h→0
f (a) exists.
is holomorphic.
Theorem 5.14 (Dunford). Let X be a Banach space, Ω ⊆ C open and T :
Proof. Let λ0 ∈ ρ(T ) and λ ∈ C with |λ − λ0 | < kR(λ0 , T )k. For fixed x ∈ X
Ω → L(X). Then the following is equivalent:
and ϕ ∈ X ′ we have by Lemma 5.7
(i) T is holomorphic in the operator norm.  X
∞  
(ii) T is strongly holomorphic. ϕ(R(λ, T )x) = ϕ (λ − λ0 )n (R(λ0 , T ))n+1 x
n=0
(iii) T is weakly holomorphic. ∞
X 
= (λ − λ0 )n ϕ (R(λ0 , T ))n+1 x
Proof. (i) =⇒ (ii) follows from the definition. (ii) ⇐⇒ (iii) follows form Theo- n=0
rem 5.13. It remains to prove (iii) =⇒ (i). As in the proof of Theorem 5.13 we
obtain for x ∈ X and ϕ ∈ X ′ where we used that the operator series converges and ϕ is continuous. Since
Z the last sum is absolutely convergent, it follows that λ → ϕ(R(λ, T )x) is ana-
1  d h ϕ(T (z)x) lytic locally at λ0 , hence holomorphic. Since weak holomorphy is equivalent to
ϕ(T (a+h)x−T (a)x) − |z=a (ϕT (z)x) = dz.
h dz 2πi Γr (z0 ) (z − a)2 (z − a − h) holomorphy in the operator norm (Theorem 5.14), the theorem is proved.
104 5.2. The resolvent Chapter 5. Spectrum of linear operators 105

The preceding theorem allows us to apply theorems of complex analysis to the Obviously xλ ∈ C 1 [0, 1], xλ (0) = 0 and x′λ (0) = λxλ (0) + y(0) = 0. Hence T − λ
resolvent map. is bijective, in particular λ ∈ ρ(T ).

Theorem 5.16. Let X be a Banach space and T ∈ L(X). Then σ(T ) 6= ∅. Note that in the last example the continuity of (T − λ) can be seen immediately:
n Z t o

Proof. Assume σ(T ) = ∅. Observe that this implies X 6= {0} and T −1 ∈ L(X). k(T − λ)−1 yk∞ = kxλ k∞ = sup eλt e−λs y(s) ds : t ∈ [0, 1]
0
Let λ ∈ C with |λ| > kT k. Then λ ∈ ρ(T ) and using the Neumann series Z 1
X ∞ X ∞
1 ≤ kyk∞ max{1, eλ } e−λs ds.

kR(λ, T )k = λn T −(n+1) ≤ |λn |kT k−(n+1) = . 0
n=0 n=0
kT k − |λ|
Definition 5.18. Let X be a Banach space The spectral radius of T ∈ L(X) is
In particular, kR(λ, T )k → 0 for |λ| → ∞. Hence for every x ∈ X and ϕ ∈ X ′
1
the map λ → ϕ(R(λ, T )x) is holomorphic and bounded in C, so constant by r(T ) := lim sup kT n k n .
the Liouville theorem. Since ϕ(R(λ, T )x) → 0 for |λ| → ∞, it follows that
ϕ(R(λ, T )x) = 0 for all λ ∈ C, x ∈ X and ϕ ∈ X ′ . By a corollary to the Theorem 5.19. Let X be a Banach space, T ∈ L(X) and r(T ) its spectral
Hahn-Banach theorem (Corollary 2.16) it follows that R(λ, T )x = 0 for all radius.
x ∈ X and λ ∈ C, hence R(λ, T ) = 0, λ ∈ C. This contradicts the fact that
(i) r(T ) ≤ kT m k1/m ≤ kT k for all m ∈ N, in particular r(T ) = lim kT mk1/m .
1 = kT T −1k ≤ kT kkT −1k = 0. m→∞
(ii) σ(T ) ⊆ {λ ∈ C : |λ| ≤ r(T )}.
The following example shows that for unbounded linear operators the cases
σ(T ) = ∅ and σ(T ) = C are possible. (iii) If X is a complex Banach space, then there exists a λ ∈ σ(T ) such that
|λ| = r(T ), in particular
Examples 5.17. (i) Let X = C([0, 1]) and r(T ) = max{|λ| : λ ∈ σ(T )}.
T : X ⊇ C 1 ([0, 1]) → X, T x = x′ .
(iv) If X is Hilbert space and T is normal, then r(T ) = kT k.
Then T is unbounded and closed and σ(T ) = σp (T ) = C.
(v) If X is a complex Hilbert space and T is normal with r(T ) = 0, then
(ii) Let X = {x ∈ C([0, 1]) : x(0) = 0}, D(T ) = {x ∈ X ∩ C 1 ([0, 1]) : x′ ∈ X} T = 0.
and
Proof. (i) Let m ∈ N arbitrary. For every n ∈ N there exist pn , qn ∈ N0 with
T : X ⊇ D(T ) → X, T x = x′ . qn < m and n = pn m + qn . Let M := max{1, kT k, . . . , kT m−1k}. Then
Then T is unbounded and closed and σ(T ) = ∅.
kT n k = kT pn m+qn k ≤ kT pn m k kT qn k ≤ M kT mkpn .
Proof. (i) Obviously, T is unbounded and densely defined. If (xn )n∈N ⊆ D(T ) 1 1 1
This implies r(T ) = lim sup kT nk n ≤ lim sup M 2 kT mk m − nm = kT m k m .
qm 1

such that xn → x and T xn → y ∈ X, then, by a theorem of Analysis 1, x is n→∞ n→∞


P∞
differentiable, hence in D(T ) and T x = x′ = y which implies that T is closed. (ii) By the formula of Hadamard, the radius of convergence of n=0 z n+1 kT nk
For every λ ∈ C the differential equation x′ −λx = 0 has the solution xλ (t) = eλt . n n
is (lim sup kT k )
1
−1 −1
= r(T ) . Hence for all λ ∈ C, |λ| > r(T ), the series
Note that xλ ∈ D(T ) and (T − λ)xλ = 0, so λ ∈ σp (T ). P∞ n→∞ −(n+1) n
n=0 λ T =: A converges in norm. By Theorem 2.10 (Neumann series),
(ii) Obviously, T is unbounded and densely defined. If (xn )n∈N ⊆ D(T ) such
A is the inverse of λ − T . Because T is closed, it follows that {λ ∈ C : |λ| >
that xn → x and T xn → y ∈ X, then, by a theorem of Analysis 1, x is
r(T )} ⊆ ρ(T ), or equivalently {λ ∈ C : |λ| ≤ r(T )} ⊆ σ(T ).
differentiable and x′ = y. Moreover, x(0) = lim xn (0) = 0, so in D(T ) and
n→∞ (iii) Let r0 := max{|λ| : λ ∈ σ(T )}. It follows from (ii) that r0 ≤ r(T ). Now
T x = x′ = y which implies that T is closed.
choose any µ ∈ C with |µ| > r0 . We have to show that |µ| > r(T ). Observe
For every λ ∈ C and every y ∈ X the initial value problem x′ − λx = y, x(0)
that by definition of R(T ) and by the formula of Hadamard
has exactly one solution xλ given by
Z t ∞
X
xλ (t) = eλt e−λs y(s) ds. (λ − T )−1 = λ−(n+1) T n , |λ| > r(T ), (5.3)
0 n=0
106 5.3. The spectrum of the adjoint operator Chapter 5. Spectrum of linear operators 107

where the series on the right hand side converges in norm. In particular, for Proof. (i) If λ ∈ σp (T ), then ker(λ − T ) ) {0}, rg(λ − T ′ ) ⊆ ker(T )◦ 6= X. It
every ϕ ∈ L(X)′ follows that λ ∈ σp (T ′ ) or λ ∈ σr (T ′ ).

(i) If λ ∈ σr (T ), then rg(λ − T ) 6= X. By Theorem 4.64 rg(λ − T ) = X if and
X only if (λ − T )′ = λ − T ′ is not injective, hence λ ∈ σp (T ′ ).
ϕ(λ − T )−1 = λ−(n+1) ϕ(T n ), |λ| > r(T ).
n=0
Theorem 5.22. Let H be a complex Hilbert space, T (H → H) a symmetric
Hence λ 7→ ϕ(T − λ)−1 defines an analytic function for |λ| > r(T ). It follows operator and λ ∈ C \ R.
from complex analysis that then the equality in (5.3) holds for all λ in the largest
open ring λ 7→ ϕ(λ − T ) is analytic, that is for all λ > r(T ). In partic- (i) k(λ − T )xk ≥ | Im(λ)| kxk for all x ∈ D(T ).
P∞ where−(n+1) In particular T − λ : D(T ) → rg(T − λ) is invertible with continuous
ular, n=0 µ ϕ(T n ) converges for every ϕ ∈ L(X)′ , hence it is weakly
convergent, and therefore (µ−(n+1) ϕ(T n ))n∈N converges to 0. It follows that inverse and the point spectrum of T is real.
(µ−(n+1) T n )n∈N is weakly convergent to 0, hence it is bounded (Corollary 3.9). (ii) If T is closed, then rg(λ − T ) is closed.
1 1 1
Let M ∈ R such that kµ−(n+1) kT n k < M , n ∈ N. Then kkT nk n < M n µ1+ n
1
for all n ∈ N, in particular r(T ) = lim kkT nk n ≤ µ. Proof. (i) For all x ∈ D(T )
n→∞
(iv) Recall that kT T ∗k = kT k2 for a normal operator T (Theorem 4.39). Hence
k(λ − T )xk kxk ≥ h(λ − T )x , xi = h(Re λ − T )x , xi + ihIm λx , xi
kT 2 k2 = kT 2 (T ∗ )2 k = k(T T ∗)2 k = k(T T ∗)k2 = kT k4, ≥ | Im λ|kxk.
n n
hence kT 2 k = kT k2 . By induction, it can be shown that hence kT 2 k = kT k2 In particular, λ − T is injective, which implies that λ ∈/ σp (T ).
for all n ∈ N, implying that (i) If (λ − T ) is continuous and closed, to its domain rg(λ − T ) is closed.
1 n 1
r(T ) = lim kT n k n = lim kT 2 k 2n = lim kT k = kT k. Theorem 5.23. Let H be a complex Hilbert space and T (H → H) a symmetric
n→∞ n→∞ n→∞
operator. Then the following is equivalent.
(v) follows directly from (iv).
(i) T is selfadjoint.
Note that in general r(T ) < kT k, for example r(T ) = 0 for every nilpotent (ii) rg(λ − T ) = H for all z ∈ C \ R.
linear operator.
(iii) rg(±i − T ) = H.
(iv) There exist z± ∈ C with Im z+ > 0 and Im z− < 0 such that rg(z± − T ) =
5.3 The spectrum of the adjoint operator H.
Lemma 5.20. (i) Let X be a Banach space and T (X → X) a densely defined (v) σ(T ) ⊆ R.
closed linear operator. Then σ(T ′ ) = σ(T ) and R(λ, T )′ = R(λ, T ′ ) for (vi) T is closed and ker(±i − T ∗ ) = H.
λ ∈ ρ(T ).
Proof. (i) =⇒ (ii) Let λ ∈ C\R. Then rg(λ−T ) 6= H is closed by Theorem 5.22
(ii) Let H be a Hilbert space and T (H → H) a densely defined closed linear
and λ∗ ∈/ σp (T ). It follows by Theorem 4.73 that
operator. Then σ(T ∗ ) = σ(T ) = {λ ∈ C : λ ∈ σ(T )} and R(λ, T )∗ =
R(λ∗ , T ∗ ) for λ ∈ ρ(T ). rg(λ − T ) = rg(λ − T )⊥⊥ = ker(λ∗ − T ∗ )⊥ = ker(λ∗ − T )⊥ = {0}⊥ = H.

Proof. The assertions follow from Theorem 4.65.


(ii) =⇒ (i) By assumption, T is symmetric, hence T ⊆ T ∗ , so it suffices to
Lemma 5.21. Let X be a Banach space and T (X → X) densely defined and show that D(T ∗ ) ⊆ D(T ). Let λ ∈ C \ R. Then λ − T and λ − T are bijective.
closed. For x ∈ D(T ∗ ) there exists a y ∈ D(T ) such that (λ − T ∗ )x = (λ − T )y. Since
(i) λ ∈ σp (T ) =⇒ λ ∈ σp (T ′ ) ∪ σr (T ′ ). T ⊆ T ∗ , it follows that T y = T ∗ , hence x − y ∈ ker(λ − T ∗ ) = {0} which implies
x = y ∈ D(T ).
(ii) λ ∈ σr (T ) =⇒ λ ∈ σp (T ′ ). (ii) =⇒ (iii) =⇒ (iv) is obvious.
108 5.3. The spectrum of the adjoint operator Chapter 5. Spectrum of linear operators 109

(iv) =⇒ (v) Let z± ∈ C with Im z+ > 0 and Im z− < 0 such that rg(z± − T ) = kxn k = 1 and kR(λn , T )xn k ≥ 12 kR(λn , T )k. From Lemma 5.7 we know
H. By Theorem 5.22, it follows that z± −T is injective and its inverse is bounded that kR(λn , T )k ≥ dist(λn1,σ(T )) . Set yn := kR(λn , T )k−1 R(λn , T )xn . Then
by |ℑz± |. Hence, by Lemma 5.7, every λ ∈ C with |λ − z± | < |ℑz± | belongs to yn ∈ D(T ) and kyn k = 1 for all n ∈ N. Moreover
ρ(T ). Given any λ ∈ C \ R, repeating the argument above finitely many times
shows that λ ∈ ρ(T ). k(λ − T )yn k ≤ k(λ − λn )yn k + k(λn − T )yn k
(v) =⇒ (ii) is obvious. = |λ − λn | + kR(λn − T )xn k−1
(vi) =⇒ (iii) Since T is closed, the range of ±i − T is closed by Theorem 5.22. ≤ |λ − λn | + 2kR(λn − T )k−1 −→ 0, n → ∞.
Therefore rg(±i − T ) = rg(±i − T )⊥⊥ = ker(∓i − T ∗ )⊥ = {0}⊥ = H.
Hence λ ∈ σap (T ).
(i) =⇒ (vi) Since T = T ∗ , it is closed and C \ R ⊆ ρ(T ), in particular ker(±i −
T ) = {0}. (iii) By Theorem 5.23 the spectrum of a selfadjoint operator is real, so σ(T ) =
∂σ(T ) ⊆ σap (T ) ⊆ σ(T ).
Analogously, we find a characterisation of essentially selfadjoint operators.
Lemma 5.27. Let H be Hilbert space and T ∈ L(H) selfadjoint. Then σ(T ) ⊆
Theorem 5.24. Let H be a complex Hilbert space and T (H → H) a symmetric [m, M ] where m := inf{hT x , xi : kxk = 1} and M := sup{hT x , xi : kxk = 1}.
operator. Then the following is equivalent. Moreover, m, M ∈ σ(T ).

(i) T is essentially selfadjoint. Proof. Let λ ∈ R, λ < m. Then λ − T is injective because for all x ∈ X
(ii) rg(λ − T ) = H for all z ∈ C \ R.
k(λ − T )xkkxk ≥ h(λ − T )x , xi ≥ (λ − m)kxk2 . (5.4)
(iii) rg(±i − T ) = H.
(iv) There exist z± ∈ C with Im z+ > 0 and Im z− < 0 such that rg(z± − T ) = In particular, rg(λ−T ) = D((λ−T )−1 ) is closed because (λ−T )−1 : rg(λ−T ) →
H. H is closed and continuous by (5.4). Hence rg(λ − T ) = rg(λ − T ) = ker(λ −
T )⊥ = H. It follows that (−∞, m) ∈ ρ(T ). Analogously (M, ∞) ∈ ρ(T ) is
(v) σ(T ) ⊆ R. shown.
(vi) ker(±i − T ∗ ) = H. Now we show that m ∈ σ(T ). By Proposition 5.26 it suffices to show that m ∈
σap (T ). By definition of m there exists a sequence (xn )n∈N such that kxn k = 1
Definition 5.25. Let X be a Banach space and T (X → X) densely defined for all n ∈ N and hT xn , xn i ց m. Since s(x, y) := h(T − m)x , yi defines a
and closed. λ ∈ C is called approximate eigenvalue if there exists a sequence positive semidefinite sesquilinear form, Cauchy-Schwarz inequality implies
(xn )n∈N ⊆ X such that kxn k = 1 for all n ∈ N and lim (T − λ)xn = 0. The set 1 1
n→∞
k(T − m)xn k2 = |s(xn , (T − m)xn )| ≤ s(xn , xn ) 2 s((T − m)xn ) 2
of all approximate eigenvalues is denoted by σap (T ). 1 1
2
= h(T − m)xn , xn i h(T − m) xn , (T − m)xn i .
2 2

Proposition 5.26. (i) Every approximate eigenvalue belongs to σ(T ).


Since the first term in the product tends to 0 for n → ∞ and the second term
3
(ii) Every boundary point of σ(T ) ⊆ C is an approximate eigenvalue of T . is bounded by (kT k − m) 2 < ∞, it follows that k(T − m)xn k tends to 0 for
n → ∞. This shows that m ∈ σap (T ). The proof of M ∈ σ(T ) is analogous.
(iii) If X is a Hilbert space and if T is selfadjoint, then every λ ∈ σ(T ) is an
approximate eigenvalue of T .
5.4 Compact operators
Proof. (i) Let λ be an approximate eigenvalue of T . Choose a sequence (xn )n∈N ⊆
D(T ) such that kxn k = 1 for all n ∈ N and (λ − T )xn → 0. Assume that Recall that a metric space M is compact if and only if every open cover of M
λ ∈ ρ(T ). Then R(λ, T ) = (λ − T )−1 is bounded, therefore contains a finite cover. M is called totally bounded if and only if every for every
ε > 0 there exists a covering of M with finitely many open balls of radius ε.
lim xn = lim R(λ − T )(λ − T )xn = R(λ − T ) lim (λ − T )xn = 0, M is called precompact (or precompact) if and only if M is compact. It can be
n→∞ n→∞ n→∞
shown that a totally bounded metric M is compact if and only if M is complete.
in contradiction to kxn k = 1 for all n ∈ N. In particular, a subset of a complete metric space is totally bounded if and only
(ii) Let λ be a boundary point of σ(T ). Then there exists a sequence (λn )n∈N ⊆ if its closure is compact. A subset of a metric space is called relatively compact
ρ(T ) which converges to λ. For every n ∈ N choose xn ∈ X such that if and only if its closure is compact.
110 5.4. Compact operators Chapter 5. Spectrum of linear operators 111

Definition 5.28. Let X, Y be normed spaces. An operator T ∈ L(X, Y ) is Theorem 5.33 (Schauder). Let X, Y be Banach space and T ∈ L(X, Y ).
called compact if for every bounded set A ⊆ X the set T (A) is relatively com- Then T is compact if and only if T ′ is compact.
pact. The set of all compact operators from X to Y is denoted by K(X, Y ).
For the proof we use the Ascoli-Arzelá theorem.
Remark 5.29. Sometimes compact operators are called completely continuous.
Theorem 5.34 (Arzelá-Ascoli). Let (M, d) be a compact metric space and
Remarks 5.30. (i) Every compact linear operator is bounded. A ⊆ C(M ) a family of real or complex valued continuous functions on M such
that
(ii) T ∈ L(X, Y ) is compact if and only if for every bounded sequence (xn )n∈N
the sequence (T xn )n∈N contains a convergent subsequence. (i) A is bounded,
(ii) A is closed,
(iii) T ∈ L(X, Y ) is compact if and only if T (BX (0, 1)) is relatively compact.
(iii) A is equicontinuous, that is,
(iv) Let T ∈ L(X, Y ) with finite dimensional rg(T ). The T is compact.
∀ε > 0 ∃δ > 0 ∀f ∈ A d(x, y) < δ =⇒ |f (x) − f (y) < ε.
(v) The identity map id ∈ L(X) is compact if and only if X is finite-dimensional.
Then A is compact.
Theorem 5.31. Let X, Y be Banach spaces. Then K(X, Y ) is a closed subspace
of L(X, Y ). Proof. See, e. g., [Rud91] or [Yos95].

Proof. Obviously, 0 ∈ K(X, Y ) and Remark 5.30 (ii) implies that the linear Proof of Theorem 5.33. First assume that T is compact. Let KX (0, 1) := {x ∈
combination of compact operators is compact. Now let (Tn )n∈N ⊆ K(X, Y ) a X : kxk < 1} be the closed unit ball in X. By assumption K := T (KX (0, 1))
Cauchy sequence. Since L(X, Y ) is complete, there exists a T ∈ L(X, Y ) such is compact in Y and bounded by kT k. Now let (ϕn )n∈N ⊆ Y ′ be a bounded
that Tn → T . We have to show T ∈ K(X, Y ). Take an arbitrary bounded sequence and M ∈ R such that kϕn k ≤ M , n ∈ N. We define the functions
sequence (xn )n∈N ⊆ X and choose M ∈ R such that kxn k ≤ M , n ∈ N.
(1) (1) fn : K → K, fn (y) := ϕn (y).
Since T1 is compact, there exists a subsequence (xn ) such that (T1 xn )n∈N
(k)
converges. Continuing like this, for every k ≥ 2 we find a subsequence (xn ) of Then (fn )n∈N is bounded by M and equicontinuous because |f (y1 ) − f (y2 )| ≤
(k−1) (k) (n)
(xn ) such that (Tk xn )n∈N converges. Let (yn )n∈N = (xn )n∈N the diagonal Cky1 −y2 k, y1 , y2 ∈ K. By the Ascoli-Arzelá, (fn )n∈N is compact, so there exists
sequence. Then, for every k ∈ N, the sequence (Tk yn )n∈N converges. Let ε > 0. a convergent subsequence (fnk )k∈N . Then also (T ′ ϕnk )k∈N converges because
ε
Choose k ∈ N such that kT −Tk k < 3M and N ∈ N such that kTk xn −Tk xm k ≤ 3ε
for m, n ≥ N . Then, for all m, n ≥ N , kT ′ ϕnk − T ′ ϕnm k = sup{kϕnk (T x) − ϕnm (T x)k : x ∈ KX (0, 1)}
= sup{kϕnk (y) − ϕnm (y)k : y ∈ K} = kfnk − fnm k.
kT yn − T ym k ≤ kT yn − Tk yn k + kTk yn − Tk ym k + kTk ym − T ymk
Mε ε Mε Now assume that T ′ is compact. Then T ′′ ∈ L(X ′′ , Y ′′ ) is compact. By
≤ + + = ε. Lemma 5.32 T ′′ ◦ JX is compact. Recall that JY ◦ T = T ◦ JX (Lemma 2.33),
3M 3 3M
so JY ◦ T : X → Y ′′ is compact. Since Y is closed in Y ′′ , T : X → Y is
Hence (T yn )n∈N is Cauchy sequence in the Banach space Y , hence convergent. compact.

Example 5.35. Let k ∈ C([0, 1]2 ) and


Lemma 5.32. Let X, Y, Z be Banach spaces, S ∈ L(X, Y ) and T ∈ L(Y, Z). Z 1
Then T S is compact if at least one of the operators S or T is compact. Tk : C([0, 1]) → C([0, 1]), (Tk x)(t) = k(s, t)x(s) ds.
0
Proof. Let (xn )n∈N be a bounded sequence in X. If S is compact, then there
exists a subsequence (xnk )k∈N such that (Sxnk )k∈N converges. By continuity of Then Tk is compact.
T , also (T Sxnk )k∈N converges. Proof. Obviously Tk is well-defined and bounded. Let (xn )n∈N ⊆ C([0, 1]) a
Now assume that T is compact. Since S is bounded, (Sxn )k∈N is bounded, bounded sequence with bound M . Hence (Tk xn )n∈N is bounded. To show that
hence there exists a subsequence (xnk )k∈N such that (T Sxnk )k∈N converges. it is equicontinuous fix ε > 0. Since k is uniformly continuous, there exists
112 5.4. Compact operators Chapter 5. Spectrum of linear operators 113

a δ > 0 such that |k(s, t) − k(s′ , t′ )| < ε if k(s, t) − (s′ , t′ )k < δ. Now for Proof. Let p := α(T ) and q := δ(T ). We divide the proof in several steps.
t1 , t2 ∈ [0, 1] with |t1 − t2 | < δ and n ∈ N we obtain Step 1. rg(T p ) ∩ ker(T n ) = {0} for every n ∈ N0 .
Z 1 To see this, choose x ∈ rg(T p ) ∩ ker(T n ). Then there exists a y ∈ X such that
|Tk xn (t1 ) − Tk xn (t2 )| ≤ |k(s, t1 ) − k(s, t2 )||xn (s)| ds < εkxn k∞ ≤ M ε. x = T p y, so 0 = T n x = T p+n y. Hence y ∈ ker T p+n = ker T p by Lemma 5.36 i.
0 It follows that x = T p y = 0.
By the Ascoli-Arzelá theorem it follows that (Tk xn )n∈N is relatively compact, Step 2. X = rg(T n ) + ker(T q ) for every n ∈ N0 .
hence it contains a convergent subsequence. For the proof fix x ∈ X. Then T q x ⊆ rg(T q ) = rg(T q+n ). Hence there exists
y ∈ X such that T q x = T q+n y. Then T q (x − T n y) = 0, and therefore x =
Let X be vector space and T : X → X a linear operator. Then obviously T n y + (x − T n y) ∈ rg(T n ) + ker(T q ).
Step 3. α(T ) ≤ δ(T ) = q.
{0} ⊆ ker T ⊆ ker T 2 ⊆ ker T 3 ⊆ . . . ,
Let x ∈ ker T q+1 . We have to show x ∈ ker T q . By step 2, with n = p,
X ⊇ rg T ⊇ rg T 2 ⊇ rg T 3 ⊇ . . . . there exist x1 ∈ rg(T p ) and x2 ∈ ker(T q ) such that x = x1 + x2 . Hence
x1 = x − x2 ⊆ ker(T q+1 ) ∩ rg(T p ) = {0} by step 1. Therefore x = x2 ∈ ker(T q ).
Lemma 5.36. Let X a vector space and T : X → X a linear operator.

(i) Assume that ker T k+1 = ker T k for some k ∈ N0 . Then ker T n = ker T k Step 4. δ(T ) ≤ α(T ) = p.
for all integer n ≥ k. By step 1 and step 2, we have that X = rg(T p ) ⊕ ker(T q ). Since rg(T p+1 ) ∩
ker(T q ) ⊆ rg(T p ) ∩ ker(T q ) = {0}, we also have X = rg(T q+1 ) ⊕ ker(T q ),
(ii) Assume that rg T k+1 = rg T k for some k ∈ N0 . Then rg T n = rg T k for implying rg R(T p+1 ) = rg(T p ), hence δ ≤ p.
all integer n ≥ k.
Theorem 5.39. Let X be a Banach space, T ∈ L(X) a compact operator and
Proof. We prove the lemma by induction. The case when n = k is clear by λ ∈ C \ {0}.
assumption.
(i) ker(λ − T )n is finite dimensional for every n ∈ N0 .
(i) Assume that n > k and ker T n = ker T k . Then
(ii) If U ⊆ X is a closed subspace with U ∩ ker(λ − T )n = {0}, then (λ − T )(U )
ker T n+1 = {x ∈ X : T n+1 x = 0} = {x ∈ X : T x ∈ ker T k } = ker T k+1 = ker T k . is closed and λ − T : U → rg((λ − T )|U has a bounded inverse.
(iii) rg(λ − T )n is closed for every n ∈ N0 .
(i) Assume that n > k and rg T n = rg T k . Then
Pn 
rg T n+1 = T (rg T n ) = T (rg T k ) = rg T k+1 = rg T k . Proof. Note that (λ − T )n = λn − n=1 nk λn−k T k and the operator sum is
compact. Hence it suffices to show the assertions for n = 1.
(i) Observe that T |ker(λ−T ) = λ id |ker(λ−T ) . Hence λ id |ker(λ−T ) is compact.
By Remark 5.30 (v) this is case if and only if ker(λ − T ) is finite dimensional.
Definition 5.37. Let X be a vector space and T : X → X a linear operator. (ii) Since U ∩ ker(λ − T ) = {0}, the restriction (λ − T )|U is invertible. We
−1
We define will show that its inverse is bounded. Assume (λ − T )|U is not bounded.
( Then there exists a sequence (xn )n∈N such that kxn k = 1 for all n ∈ N and
min{k ∈ N0 : ker T k = ker T k+1 }, if the minimum exists, lim (λ − T )xn = 0. Since T is compact, there exists a convergent subsequence
ascent of T := α(T ) := n→∞
∞ else (T xnk )k∈N . Hence
(
k k+1
min{k ∈ N0 : rg T = rg T }, if the minimum exists,
descent of T := δ(T ) := λxnk = T xnk + (λ − T )xnk −→ lim T xnk =: y.
∞ else. | {z } n→∞
→0

Lemma 5.38. Let X be a vector space and T : X → X a linear operator. If Note that y ∈ U because U is closed. Moreover, y ∈ ker(λ − T ) because
both the ascent α(T ) and the descent δ(T ) are finite, then α(T ) = δ(T ) =: p
and X = rg(T p ) ⊕ ker(T p ). (λ − T )y = (λ − T ) lim xnk = lim (λ − T )xnk = 0.
n→∞ n→∞
114 5.4. Compact operators Chapter 5. Spectrum of linear operators 115

Hence y ∈ ker(λ − T ) ∩ U = {0} in contradiction to kyk = lim kλxn k = λ 6= 0. Therefore (T xn )n∈N does not contain a convergent subsequence in contradiction
n→∞
−1 to T being compact.
Hence (λ − T )|U : rg(λ − T )|U → U is bounded. Since it is also closed, its
domain rg(λ − T )|U must be closed.
Theorem 5.42 (Spectrum of a compact operator). Let X be a Banach
(iii) By (i) we already know that dim ker(λ − T ) < ∞. Then by the following space. For a compact operator T ∈ L(X) the following holds.
lemma 5.40 there exists a closed subspace U ⊆ X such that X = ker(λ − T ) ⊕ U .
Hence rg(λ − T ) = rg((λ − T )|U ) is closed by (ii). (i) If λ ∈ C \ {0}, then λ either belongs to ρ(T ) or it is an eigenvalue of T ,
that is C \ {0} ⊆ ρ(T ) ∪ σp (T ).
Lemma 5.40. Let X be a Banach space and M ⊆ X a finite dimensional
subspace. Then there exists a closed subspace U of X such that X = M ⊕ U . (ii) The spectrum of T is at most countable and 0 is the only possible accu-
mulation point.
Proof. Let x1 , . . . , xn a basis of M . Then there exist ϕ1 , . . . , ϕn ∈ M ′ such that (iii) If λ ∈ σ(T ) \ {0}, then the dimension of the algebraic eigenspace Aλ (T )
kϕk k = 1 and ϕk (xj )δkj for all j, k = 1, . . . , n. By the Hahn-Banach theorem is finite and Aλ (T ) = ker(λ − T )p where p is the Riesz index of λ − T .

the ϕk can be extended P to functionals ψk ∈ X with kψk k = 1, k = 1, . . . , n. Let
P : X → X, P x = nj=1 ϕj (x)x. Obviously P = P 2 , hence P is a projection. (iv) X = ker(λ − T )p ⊕ rg(λ − T )p for λ ∈ σ(T ) \ {0} where p is the Riesz
Note that M = P (X). Hence X = rg(P ) ⊕ ker P = M ⊕ ker P . index of λ − T and ker(λ − T )p and rg(λ − T )p are T -invariant.

Theorem 5.41. Let X be a Banach space, T ∈ L(X) a compact operator and (v) σp (T ) \ {0} = σp (T ′ ) \ {0} and σ(T ) = σ(T ′ ). If H is a Hilbert space then
λ ∈ C\{0}. Then α(λ−T ) = δ(λ−T ) = p < ∞ and X = ker(λ−T )p ⊕rg(λ−T )p . σp (T ) \ {0} = {λ ∈ C : λ ∈ σp (T ∗ )} \ {0} = σp (T ∗ ) \ {0}, where the bar
denotes complex conjugation, and σ(T ) = {λ ∈ C : λ ∈ σ(T ∗ )} = σ(T ∗ ).
The number p = α(λ − T ) = δ(λ − T ) is called the Riesz index of λ − T .
Proof. (i) Let λ ∈ C \ {0}. By Theorem 5.41 the Riesz index p of λ − T is
Proof. By Lemma 5.38 it suffices to show that α(T ) and δ(T ) are finite. finite. If p = 0, then X = rg(λ − T ) by the proof of Lemma 5.38 (step 2), hence
Assume that α is not finite. Since in this case ker(λ − T ) ( ker(λ − T )2 ( . . . λ ∈ ρ(T ). If p 6= 0, then λ ∈ σp (T ).
we can find a sequence (xn )n∈N ⊆ X such that for all n ∈ N
(ii) It suffices to show that for every ε > 0 the set {λ ∈ σ(T ) : |λ| > ε} is finite.
1 Assume there exists an ε > 0 such that the set is not finite. Then there exists
kxn k = 1, xn ∈ ker(λ − T )n , and kxn − zk ≥ for all z ∈ ker(λ − T )n−1 .
2 a sequence (λn )n∈N such that λn 6= λm for n 6= m and |λn | > ε, n ∈ N. Since
σ(T ) \ {0} consists of eigenvalues, we can choose eigenvectors xn of T with
The last condition can be satisfied by the Riesz lemma (Theorem 1.21) because
eigenvalues λn . Note that the xn are linearly independent because λn 6= λm
ker(λ − T )n is closed for all n ∈ N. Then for all 1 ≤ m < n
for n 6= m. Let Un := span{x1 , . . . , xn }. Note that all Un are T -invariant,
1 closed and that U1 ( U2 ( U3 ( . . . . Using the Riesz Lemma, we can choose a
kT xn − T xm k = kλxn −λxm − (λ − T )xn + (λ − T )xm k ≥ .
| {z } 2 sequence (yn )n∈N such that for all n ∈ N
∈ker(λ−T )n−1
1
Therefore (T xn )n∈N does not contain a convergent subsequence in contradiction kyn k = 1, yn ∈ Un , and kyn − zk ≥ for all z ∈ Un−1 .
2
to T being compact. P
Assume that δ is not finite. Since in this case rg(λ − T ) ) rg(λ − T )2 ) . . . we Let 1 ≤ m < n. Note that T ym ∈ Um . Let yn = nj=1 αj xj for some αj ∈ C.
can choose a sequence (xn )n∈N ⊆ X such that for all n ∈ N Then
n
X n
X
1
kxn k = 1, xn ∈ rg(λ − T )n , and kxn − zk ≥ for all z ∈ rg(λ − T )n+1 . (λn − T )yn = αn (λn − T )xn + αj (T − λn )xj = αj (λj − λn )xj ∈ Un−1 .
2
j=1 j=1
n
The last condition can be satisfied by the Riesz lemma because rg(λ − T ) is
closed for all n ∈ N by Theorem 5.39. Then for all 1 ≤ m < n Hence
1 1
kT xn − T xm k = kλxn −λxm − (λ − T )xn + (λ − T )xm k ≥ . kT yn − T ym k = kλn yn −(λn − T )yn − T ym k ≥ . (5.5)
| {z } 2 | {z } 2
∈rg(λ−T )n+1 ∈Un−1
116 5.4. Compact operators Chapter 5. Spectrum of linear operators 117

Therefore (T xn )n∈N does not contain a convergent which contradicts the as- (b) (B) and (D) have non-trivial solutions. In this case dim(ker(λ−T )) =
sumption that T is compact. dim(ker(λ − T ′ )) > 0 and (A) and (C) have solutions if and only if
(iii) and (iv) follow from Theorem 5.42. ϕ(y) = 0 for all solutions ϕ of (D),
(v) By Schauder’s theorem T ′ is compact (theorem 5.33) Hence for λ ∈ C it η(x) = 0 for all solutions x of (B).
follows that
Definition 5.45. Let X, Y be Banach spaces. T ∈ L(X) is called Fred-
λ ∈ ρ(T ) ⇐⇒ ker(λ − T ) = {0} and rg(λ − T ) = X holm operator if rg(T ) is closed and n(T ) := dim(ker T ) < ∞ and d(T ) :=

⇐⇒ rg(λ − T ′ ) = {0} and ◦ ker(λ − T ′ ) = X codimY (rg T ) := dim(Y / rg(T )) < ∞. In this case, χ(T ) := n(T ) − d(T ) is
⇐⇒ rg(λ − T ′ ) = X ′ and ker(λ − T ′ ) = {0} called the Fredholm index.
⇐⇒ λ ∈ ρ(T )
Proof of Theorem 5.44. . . . . . . . . . . . .
Theorem 5.43 (Fredholm alternative; Riesz-Schauder theory). Let X Now we return to the spectrum of compact operators.
be a Banach space, T ∈ L(X) a compact operator and λ ∈ C \ {0}. Then exactly
one of the following is true: Lemma 5.46. Let H be Hilbert space, 6= {0}, and T ∈ L(H) a selfadjoint
compact operator. Then at least one the values kT k or −kT k is an eigenvalue
(i) For every y ∈ X the equation (λ−T )x = y has exactly one solution x ∈ X.
of T . In particular, if T 6= 0, then T has at least one eigenvalue distinct from
(ii) (λ − T )x = 0 has a non-trivial solution x ∈ X. 0.

Proof. (i) is equivalent to λ ∈ ρ(T ) and (ii) is equivalent to λ ∈ σp (T ). Since λ 6= Proof. If kT k = 0, the assertion is clear. Now assume that kT k 6= 0. Recall
0, the latter is equivalent to λ ∈ σ(T ). The assertion follows from Theorem 5.42. that kT k = sup{|hT x , xi| : x ∈ X, kxk = 1} (Theorem 4.45).
By Lemma 5.27 the numbers m = inf{hT x , xi : x ∈ X, kxk = 1} and M =
inf{hT x , xi : x ∈ X, kxk = 1} belong to the spectrum of T . Since T is compact
A more precise formulation of the Fredholm alternative is the following. and kT k 6= 0, it follows that ∅ 6= {±kT k} ∩ σ(T ) = {±kT k} ∩ σp (T ).

Theorem 5.44. Let X be a Banach space, T ∈ L(X) a compact operator and Theorem 5.47 (Spectral theorem for compact selfadjoint operators).
λ ∈ C \ {0}. For x, y, ∈ X and ϕ, η ∈ X ′ consider the equations Let H be a Hilbert space and T ∈ L(H) a compact selfadjoint operator.
(A) (λ − T )x = y, (C) (λ − T ′ )ϕ = η, (i) There exists an orthonormal system (en )N
n=1 of eigenvectors of T with
(B) (λ − T )x = 0, (D) ′
(λ − T )ϕ = 0. eigenvalues (λn )N
n=1 where N ∈ N ∪ {∞} such that

N
X
Then Tx = λn hx , en i en , x ∈ H. (5.6)
n=1
(i) For y ∈ X the following is equivalent:
The λn can be chosen such that |λ1 | ≥ |λ2 | ≥ · · · > 0. The only possible
(a) (A) has a solution x. accumulation point of the sequence (λn )n∈N is 0.
(b) ϕ(y) = 0 for every solution ϕ of (D).
(ii) If P0 is the orthogonal projection on ker T , then
(ii) For η ∈ X ′ the following is equivalent: N
X
x = P0 x + hx , en i en , x ∈ H. (5.7)
(a) (C) has a solution ϕ.
n=1
(b) η(x) = 0 for every solution x of (B).
(iii) If λ ∈ ρ(T ), λ 6= 0
(iii) Fredholm alternative: Exactly one of the following holds:
XN
hx , en i
(a) For all y ∈ X and η ∈ X ′ the equations (A) and (C) have exactly one (λ − T )−1 x = λ−1 P0 x + en , x ∈ H.
λ −λ
n=1 n
solution (in particular (B) and (D) have only the trivial solutions).
118 5.4. Compact operators Chapter 5. Spectrum of linear operators 119

Proof. (i) Let X1 = X and T1 = T . If T 6= 0, then there exists a λ1 ∈ σp (T1 ) The representation (5.8) allows us to define the root of a positive compact
such that |λ1 | = kT1 k 6= 0. Let B1 be an orthonormal basis of ker(λ1 −T1 ). Note selfadjoint operator.
that B1 is finite because T is compact (Theorem 5.42). Let X1 := ker(λ1 −T )⊥ =
rg(λ1 − T ) = rg(λ1 − T ). Here we used that T is selfadjoint and consequently Theorem 5.49. Let H be a Hilbert space and K ∈ L(H) a compact operator.
λ ∈ σp (T ) ⊆ R. By Theorem 5.42, X2 is T1 -invariant, hence T2 := T1 |X2 ∈
(i) T is positive ⇐⇒ all eigenvalues of T are positive.
L(X2 ). Obviously, T2 is selfadjoint and compact. If T2 6= 0, then there exists
a λ2 ∈ σp (T2 ) such that |λ2 | = kT2 k = 6 0. Let B2 be an orthonormal basis T is strictly positive ⇐⇒ all eigenvalues of T are strictly positive.
of ker(λ2 − T2 ). Note that B1 is finite because T is compact (Theorem 5.42). (ii) If T is positive and k ∈ N then there exists exactly one positive compact
Hence B1 ∪ B2 is an orthonormal basis of span{ker(λ1 − T ), ker(λ2 − T )}. Let selfadjoint operator R such that Rk = T .
X3 := span{ker(λ1 − T ), ker(λ2 − T )}⊥ and T3 := T2 |X3 . Continuing like this we
obtain a sequence of Banach spaces Xn and a sequence of compact selfadjoint Note that the theorem does not imply that there cannot be non-compact oper-
operators Tn ∈ L(Xn ). Let x ∈ X. Define ators A ∈ L(H) such that A2 = T . In Corollary 5.60 we will show that every
X bounded positive selfadjoint operator has a unique positive root.
xn+1 = x − hx , en i en ∈ Xn+1 .
en ∈B1 ∪...Bn Proof of Theorem 5.49. Recall that a linear operator T is positive if and only if
It follows that hT x , xi ≥ 0 for all x ∈ H. Let P0 , λn and en as in (5.7). Then (i) follows from
X DX E X
kT x − T hx , en i en k = kTn+1 xn+1 k ≤ |λn+1 |kxk −→ 0, n → ∞. X
hT x , xi = λn hx , en i en , P0 x + λn hx, en , ein = λn |hx , en i|2 ≥ 0.
en ∈B1 ∪...Bn
n n n
This implies that P 1/k k
N
X N
X
For the proof of (ii) define R = n λn h· , en i en . Obviously R = T . To
Tx = hx , en iT en = λn hx , en i en . show uniqueness, assume that there exists a compact selfadjoint positive linear
k
n=1 n=1 operator
P S such that S = T . Since S is compact, it has a representation
S = n µn Qn with pairwise orthogonal projections Qn . By assumption
(ii) Note that . . .
X
(iii) T = Sk = µkn Qn .
n

Corollary 5.48. Let H be a Hilbert space and T ∈ L(H) a compact selfadjoint Hence the µn are the kth roots the eigenvalues λn of T , so S = R.
operator. There exists a sequence (Pn )N
n=1 of pairwise orthogonal projections
with N ∈ N ∪ {∞} and a sequence |λ1 | ≥ |λ2 | ≥ . . . such that Definition 5.50. Let H be a Hilbert space and T ∈ L(H) a positive selfadjoint
1
compact operator. Then |T | := (T ∗ T ) 2 . The non-zero eigenvalues sn of |T | are
N
X the singular values of T .
T = λn Pn (5.8)
n=1
Obviously |T | and |T ∗ | are positive selfadjoint compact operators.
where the series converges to T in the operator norm. If (λn )n is an infinite
sequence, then lim λn = 0. The representation (5.8) is unique if the λn are Lemma 5.51. (i) k |T |x k = kT xk and k |T ∗|y k = kT ∗ yk and for x ∈ H1
n→∞
pairwise distinct. and y ∈ H2 .

Proof. If the series is a finite sum, the assertion is clear. Now assume (ii) s is a singular value of T if and only if s2 is an eigenvalue of T ∗ T and
P∞ that the T T ∗.
series is an infinite. Note that for every k ∈ N the operator n=k λn Pn is
normal and that the norm of a normal operator is equal to maximum of the
moduli of the elements of its spectrum (Theorem 5.19). Since |λk+1 | → 0 for Proof. (i) For all x ∈ H1
k → ∞ the claim follows from k |T |x k2 = h|T |x , |T |xi = h|T |2 x , xi = hT ∗ T x , xi = kT xk2 .
Xk

T − λn Pn = sup{|λn | : n ≥ k + 1} = |λk+1 |. An analogous calculation shows k |T ∗ |y k = kT ∗yk and for y ∈ H2 .
n=1 (ii) follows from the uniqueness of the representation (5.8).
120 5.4. Compact operators Chapter 5. Spectrum of linear operators 121

Note that |T | can be defined more generally for positive selfadjoint operators 5.5 Hilbert-Schmidt operators
on a Hilbert space H, see Definition 5.61.
A representation similar to (5.6) exists for arbitrary compact operators. Definition 5.54. Let H1 , H2 be Hilbert spaces and K ∈ L(H1 , H2 ). K is called
a Hilbert-Schmidt operator if and only if there exists an ONB (eλ )λ∈Λ of H1 such
Theorem 5.52. Let H1 , H2 be Hilbert spaces and T ∈ L(H1 , H2 ) a compact that
operator. X
kK eλ k2 < ∞.
(i) Let s1 ≥ s2 ≥ · · · > 0 be the singular values of T and (ϕn )N
n=1 ⊆ H1 and λ∈Λ
(ψn )N
n=1 ⊆ H2 such that
The set of all Hilbert-Schmidt operators from H1 to H2 is denoted by HS(H1 , H2 ).
N
X
Tx = sn hx , ϕn iψn , x ∈ H1 , Theorem 5.55. Let H1 , H2 be Hilbert spaces.
n=1
N
X (i) A operator K ∈ L(H1 , H2 ) is a Hilbert-Schmidt operator if and only if K ∗
T ∗y = sn hy , ψn iϕn , y ∈ H2 . is a Hilbert-Schmidt operator. In this case:
n=1
X X X
If there are infinitely many sn , then lim sn = 0. kK eα k2 = kK eβ k2 = kK eλ k2 < ∞
n→∞ α∈A β∈B λ∈Λ
(ii) The non-zero eigenvalues of |T | and |T ∗ | coincide and are equal to the sn .
The s2n are the eigenvalues of T ∗ T and T T ∗. Moreover, the ψn = s1n T ϕn for all ONBes (eα )α∈A of H1 and (eβ )β∈B of H2 .
are eigenvectors of T ∗ .
(ii) Every Hilbert-Schmidt operator is compact.
Proof. (i) Let (ϕn )n∈N ⊆ H1 a ONS such that, see Theorem 5.47,
(iii) Let K ∈ L(H1 , H2 ) be a compact operator with singular values x1 ≥ x2 ≥
N
X N
X s3 ≥ . . . . Then K is a Hilbert-Schmidt operator if and only if K ∗ is a

|T |x = sn hx , ϕn iϕn , T Tx = s2n hx , ϕn iϕn . Hilbert-Schmidt operator if and only if
n=1 n=1 X
1 s2n < ∞.
Let ψn := xn T ϕn . Then (ψn )n∈N is an ONS in H2 because
n
1 1 1
hψn , ψm i = 2 hT ϕn , T ϕm i = 2 hT ∗ T ϕn , ϕm i = 2 s2n δnm = δnm . Theorem 5.55 (i) shows that for K ∈ HS(H1 , H2 ) the Hilbert-Schmidt norm
sn sn sn
Moreover X  21
kKkHS := kK eα k2 for an ONB (eα )α∈A .
1 s2n
T T ∗ ψn = T T ∗ T ϕn = T ϕn = s2n ψn . α∈A
sn sn
is well-defined.
Hence σp (T ∗ T )\{0} = {s2n : 1 ≤ n ≤ N } ⊆ σp (T T ∗ )\{0}. Similarly the reverse
inclusion can be shown, so that σp (T ∗ T ) \ {0} ⊆ σp (T T ∗ ) \ {0}. Proof of Theorem 5.55. (i)P Let K be a Hilbert-Schmidt operator and (eλ )λ∈Λ
(ii) . . . an ONB of H1 such that λ∈Λ kK eλ k2 < ∞. For an arbritray ONB (ψβ )β∈B
of H2 we find, using Parseval’s equality (Theorem 4.31) First we show that K ∗
Theorem 5.53 (Min-Max-Principle). Let H1 , H2 be Hilbert spaces, K ∈ is also a Hilbert-Schmidt operator.
L(H1 , H2 ) a compact operator with singular values s1 ≥ s2 ≥ s3 ≥ . . . . Then 2
s1 = kKk and for n ≥ 2 X X X ∗

XX
∗ 2
n o kK ψβ k = hK ψβ , eλ i eλ = |hK ∗ ψβ , eλ i|2

sn+1 = inf sup kKxk : x ∈ H1 , x ⊥ span{x1 , . . . , xn }, kxk = 1 . β∈B β∈B λ∈Λ λ∈Λ β∈B
x1 ,...xn ∈H1 XX X
= |hψβ , K eλ i|2 = kKψλ k2 < ∞.
Proof. . . . λ∈Λ β∈B λ∈Λ
122 5.5. Hilbert-Schmidt operators Chapter 5. Spectrum of linear operators 123

In particular, the Hilbert-Schmidt norm of K ∗ does not depend on the chosen (i) T is a Hilbert-Schmidt operator.
ONB of H2 . Applying the same proof to K ∗ , it follows that the Hilbert-Schmidt
norm of K = K ∗∗ does not depend on the chosen ONB of H1 . For the proof of (ii) There exists a k ∈ L2 (0, 1)2 such that
kKk ≤ kKkHS we observe that every x ∈ H1 with kxk = 1 can be extended to Z 1
a ONB of H1 . Hence (T x)(t) = k(s, t)x(s) ds.
0
kKkHS ≥ kKxk ≥ sup{kKyk : y ∈ H1 , kyk = 1} = kKk.
In this case we write Tk for T .
(ii) Let (eλ )λ∈Λ an ONS of H1 and (en )n∈N a subset containing all eλ with If one of the equivalent conditions holds, then
K eλ 6= 0 (this family is at most countable by Lemma 4.27). For n ∈ N let Pn be Z 1Z 1 2
the orthogonal projection on {e1 , . . . , en }. Note that all Pn are compact because kT k = |k(s, t)|2 ds. dt = kkkL2(0,1)2 .
they have finite-dimensional range. Since K is a Hilbert-Schmidt operator, we 0 0

find that
Proof. (ii) =⇒ (i) Let (en )n be an ONB of L2 (0, 1). Then also (en )n is an

X ONB of L2 (0, 1) (where en denotes the to en complex conjugated function) and
kK − KPn k2 = kK(id −Pn )k2 ≤ kK(id −Pn )k2HS = kK en k2 −→ ∞, we find
m=n+1

X X∞ Z 1 Z 1 2 ∞ Z 1
X

in particular K is compact because it is the norm limit of compact operators. kT en k2 = k(s, t) en (s) ds dt = |hk(·, t) , en i|2 dt

(iii) Assume that K is compact. By Theorem 5.52 n=1 n=1 0 0 n=1 0
P we can choose ONSs Z 1X∞
(ϕn )λ∈N of H1 and (ψn )λ∈N of H2 such that Kx = n∈N sn hx , ϕn iψn where 2
= |hk(·, t) , en i| dt (5.9)
s1 ≥ s2 ≥ · · · ≥ 0 are the singular values of K. 0 n=1
If K is a Hilbert-Schmidt operator, then Z 1
= kk(·, t)k2 dt (5.10)
N
X N
X 0
s2n = 2
kKϕn k ≤ kKk2HS < ∞. Z 1Z 1
n=1 n=1 = |k(s, t)|2 ds dt = kkkL2 (0,1)2 .
0 0
PN
Now assume that n=1 s2n < ∞ and choose an arbitrary ONB of H1 containing In (5.9) we have used the monotone convergence theorem to exchange the
(ϕn )n∈N . It follows that sum and the integral (Theorem A.18) and in (5.10) we used Parseval’s equal-
N N ity (Theorem 4.31). It follows that T is a Hilbert-Schmidt operator and that
X X X
kKϕλ k2 = kKϕn k2 ≤ kKk2HS = s2n < ∞, kT kHS = kkkL2 (0,1)2 .
λ∈Λ n=1 n=1 (i) =⇒ (ii) By the proof we have an isometry
implying that K is a Hilbert-Schmidt operator. Ψ : L2 (0, 1)2 → HS(L2 (0, 1)), Ψk = Tk .

Lemma 5.56. The finite-rank operators are dense in the Hilbert-Schmidt op- We will show that the range of Ψ is dense in HS(H). By Lemma 5.56 it suffices
erators. to show that rg(Ψ) containsPthe finite-rank operators. Let T be of finite rank.
n0
Then T is of the form T = n=1 h· , xn iyn so that for every f ∈ H
Proof. Let H be a Hilbert space and S ∈ HS(H). In particular, S is compact n
X n
X0 Z 1 Z 1X n0 
P 0

and there exist ONBs (ϕn )n∈N and (ψn )n∈N such that S = Nn=1 sn h· , ϕn iψn . T f (t) = hf , xn iyn (t) = f (s)xn (s)yn (t) ds = xn (s)yn (t) f (s) ds.
PM 0 0
For M ∈ N let us define SM = n=1 sn h· , ϕn iψn . Then kS − SM k2 ≤ kS − n=1 n=1 n=1
P N
SM k2HS = n=M+1 s2n → 0 for M → ∞. This shows that T ∈ rg Ψ. Fix S ∈ HS(H) and choose a sequence (Sn )n∈N in
the range of Ψ. Since Ψn is an isometry, it follows that (Ψ−1 Sn )n∈N is Cauchy
An important class of examples is given in the following theorem.
sequence in H, hence its limit exists. Using the continuity of Ψ we find
 
Theorem 5.57. Let H = L2 (0, 1) and T ∈ L(H). Then the following is equiv-
S = lim Sn = lim ΨΨ−1 Sn = Ψ lim Ψ−1 Sn ∈ rg(Ψ).
alent: n→∞ n→∞ n→∞
124 5.6. Polar decomposition Chapter 5. Spectrum of linear operators 125

Theorem 5.58. Let H1 , H2 be Hilbert spaces. Proof. Without restriction we can assume kT k ≤ 1, hence 0 ≤ T ≤ id. Now
  assume that a solution R ∈ L(H) of R2 = T exists. Let A := id −T and
(i) HS(H1 , H2 ), k · kHS is a normed spaces. The norm is induced by the X := id −R. Note that
inner product
id −A = T = R2 = (1 − X)2 = id −2X + X 2 .
X
hS , T iHS = hS eα , T eα i, S, T ∈ HS(H1 , H2 ),
α Note that 0 ≤ R ≤ id if and only if 0 ≤ X ≤ id. Hence R is a non-negative
solution of R2 = T if and only if X is a non-negative solution of
for an arbitrary ONB (eα )α∈A of H1 .
1
X= (A + X 2 ). (5.11)
(ii) Let T ∈ HS(H1 , H2 ) and A a bounded linear operator between appropriate 2
Hilbert spaces. Then AT and T A are Hilbert-Schmidt operators and
Step 1. Construction of a solution of (5.11).
kAT kHS ≤ kAk kT kHS , kT AkHS ≤ kAk kT kHS . We define
1 2
(iii) HS(H) is a two-sided ideal in L(H). X0 := id, Xn := (A + Xn−1 ), n ∈ N.
2
Proof. Note that (a+b)2 = a2 +b2 +2ab = a2 +b2 −(a−b)2 +a2 +b2 ≤ 2(a2 +b2 ) Note that every Xn is a polynomial in A with positive coefficients and that
for a, b ∈ R. Xn Xm = Xm Xn for all n, m ∈ N. Since A is positive, this implies that all Xn
(i) Let S, T ∈ HS(H1 , H2 ) and λ ∈ C. Then obviously λS ∈ HS(H1 , H2 ). To are positive. We will show the following properties of the sequence (Xn )n∈N by
show that S + T ∈ HS(H1 , H2 ) fix an ONS (eλ )λ∈Λ of H1 . Using the above induction.
remark is follows that
X X X (i) Xn − Xn−1 is a polynomial in A with positive coefficients, so that in
k(S + T ) eλ k2 ≤ (kS eλ k + kT eλ k)2 ≤ 2 kS eλ k2 + kT eλ k2 < ∞. particular Xn − Xn−1 ≥ 0.
λ∈Λ λ∈Λ λ∈Λ
(ii) kXn k ≤ 1.
It follows that h· , ·iHS is well-defined. The properties of an inner product are
clear. In particular, kT kHS = hT , T iHS for T ∈ HS(H1 , H2 ). All assertions are clear in the case n = 0 (with X−1 := 0). Now assume that
the assertions are true for some n ∈ N. Note that
(ii) Note that
X X 1 1 1
Xn+1 − Xn = (A + Xn2 ) − (A + Xn−1
2
) = (Xn2 − Xn−1
2
)
kAT eλ k2 ≤ kAk2 kT eλ k2 = kAk2 kT k2HS , 2 2 2
λ∈Λ λ∈Λ 1
= (Xn − Xn−1 )(Xn + Xn−1 ).
2
so AT is a Hilbert-Schmidt operator. It follows that T A = (A∗ T ∗ )∗ is also a
Hilbert-Schmidt operator with norm kT AkHS = k(A∗ T ∗ )∗ kHS = kA∗ T ∗ kHS ≤ Since by induction hypothesis both terms in the second line are polynomials in
kA∗ k kT ∗kHS ≤= kAk kT kHS . A with positive coefficients, (i) is proved for n + 1. (ii) follows from kXn+1 k ≤
1
(iii) is a consequence of (i) and (ii). 2 (kAk + kXn+1 k) ≤ 1.
Since (Xn )n∈N is uniformely bounded monotonically increasing sequence in,
there exists an X ∈ L(H) such that X = s- lim Xn and kXk ≤ lim inf n→∞ kXn k ≤
n→∞
5.6 Polar decomposition 1 (see Exercise 4.25).
Theorem 5.59. Let H be a Hilbert space and T ∈ L(H) a selfadjoint operator Now let S ∈ L(H) with ST = T S. By definition of A, then also SA = AS and
with T ≥ 0. Then there exists exactly one R ∈ L(H) such that R ≥ 0 R2 = T . Xn S = Xn S for all Xn since the Xn are polynomials in A. For every x ∈ H we
therefore obtain
In addition, if S ∈ L(H) commutes with T , then S commutes with R.
√ 0 ≤ kSXx − XSxk = lim kSXn x − Xn Sxk = lim kSXn x − SXn xk = 0.
the operator R is called the root of T and is denoted by T. n→∞ n→∞
126 5.6. Polar decomposition Chapter 5. Spectrum of linear operators 127

Since all Xn commute with T , it follows that Xn X = Xn X for all n ∈ N, so Theorem 5.63 (Polar decomposition). Let H1 , H2 be Hilbert spaces and
that for all x ∈ X T ∈ L(H1 , H2 ). Then there exists a partial isometry U ∈ L(H1 , H2 ) such that
T = U |T |. If in addition the initial space of U is (ker T )⊥ , then U is unique.
k(Xn2 − X 2 )xk = k(Xn − X)(Xn + X)xk ≤ 2k(Xn − X)2 xk −→ 0, n → ∞,
2
Proof. Note that k |T |x k2 = kT xk2 for all x ∈ H1 because
which shows that X = s- lim Xn2 . Therefore X solves (5.11) because
n→∞ 1 1
k |T |x k2 = h|T |x , |T |xi = h(T ∗ T ) 2 x , (T ∗ T ) 2 xi = hT ∗ T x , xi = hT x , T xi = kT xk2 .
1 1 1
X = s- lim Xn = s- lim (A + Xn2 ) = (A + s- lim Xn2 ) = (A + X 2 ).
n→∞ n → ∞2 2 n→∞ 2 We define

Setting R = id −X we obtain a bounded selfadjoint solution of R2 = T with U : rg(|T |) → rg(T ), U (|T |x) = T x.
0 ≤ R ≤ id.
U is well-defined because for x, y ∈ H1 with |T |x = |T |y it follows that kT x −
Step 2. Uniqueness of the solution.
T yk = k|T |x − |T |yk = 0 hence T x = T y. U is and isometry because kT xk =
Let R′ ∈ L(H) be solution of R2 = T with R′ ≥ 0. Then R and R′ commute k |T |x k for all x ∈ H as shown above. In particular, kU k = 1 and has a unique
because continuous extension to rg(|T |) → rg(T ). Now we extend U to H1 by setting

R′ A = R′ (R′ )2 = (R′ )2 R′ = AR′ . U x = 0 for all x ∈ rg(|T |) = ker(|T |) = ker T .

It follows that

(R − R′ )R(R − R′ ) + (R − R′ )R′ (R − R′ ) = (R2 − R′2 )(R − R′ ) = 0.

Since both operators on the left hand side are non-negative, it follows that both
of them are 0 and therefore

(R − R′ )4 = (R − R′ )R(R − R′ ) − (R − R′ )R′ (R − R′ ) = 0.

Since R − R′ is normal, it follows that k(R − R′ )k4 = k(R − R′ )k4 .

Corollary 5.60. If S, T ∈ L(H) are positive and ST = T S, then also ST is


positive.

Proof. By Theorem 5.59 the root of T exists, is selfadjoiunt and commutes with
S. Hence for all x ∈ H
√ √ √ √ √ √
hST x , xi = hS T T x , xi = h T S T x , xi = hS T x , T xi ≥ 0.
1
Definition 5.61. For T ∈ L(H) we define |T | := (T ∗ T ) 2 .

Definition 5.62. Let H1 , H2 be Hilbert spaces and U ∈ L(H1 , H2 ). U is called


a partial isometry if U |(ker U)⊥ is an isometry. ker U ⊥ is called its initial space.

Note that U is an partial isometry if and only if

U |(ker U)⊥ : (ker U )⊥ → rg(U )

is unitary.
Chapter A. Lp spaces 129

Appendix A

Lp spaces

Spaces of integrable functions play an important role in applications. As the


norm of a function f : [a.b] → K one could consider
Z b
kf k := |f (t)| dt,
a

or more generally, for some 1 ≤ p < ∞


Z ! p1
b
kf kp := |f (t)|p dt .
a

It can be shown that k·kp is a norm on C([a, ]). However the space of continuous
functions C([a, b]) is not complete for the norm k · k1 . For example, let
(
tn , 0 ≤ t ≤ 1,
fn : [0, 2] → R, fn (t) =
1, 1 < t ≤ 2.

All fn are continuous and it is easy to check that kfn − fm k1 → 0 for n, m → ∞.


So the fn form a Cauchy sequence, but it is not convergent. (If it were, then
there must exist a continuous function g such that
Z 2 Z 1 Z 2
|fn (t) − g(t)| dt = |fn (t) − g(t)| dt + |fn (t) − g(t)| dt −→ 0
0 0 1

for n → ∞. Hence g(t) = 0 for t ∈ (0, 1) and g(t) = 0 for t ∈ (1, 2) which is
impossible for a continuous function by the intermediate value theorem.
If we extend the space of functions to the Riemann integrable functions R([a, b]),
then the sequence above does converge to χ[1,2] . But there are several other
problems with the space of Riemann integrabel functions.
For example, let Q∩[0, 1] = {qn : n ∈ N}. Then all characteristic functions χn :=
χ{q1 , ...,qn } are Riemann integrable, kχn k1 = 0, the χn form a Cauchy sequence,
130 A.1. A reminder on measure theory Chapter A. Lp spaces 131

the pointwise limit exists and is χQ∩[0,1] which is not Riemann integrabel. This Example A.3. The smallest σ-algebra containing all intervals of R is called
example shows that k·k1 is only a seminorm on R([a, b]), that Cauchy sequences the Borel sets.
do not need to converge and that in general pointwise limit and integral cannot More generally, let (T, O) be a topological space. Then the Borel sets is the
be exchanged. The pointwise limit of a sequence of Riemann integrable functions σ-algebra generated by O.
does not need to be Riemann integrable.
Recall that the Riemann integral of a function f : [a, b] → R is obtained as The aim is to assign a measure µ(U ) to every Borel set U ⊆ R such that the
the limit of Riemann sums when the interval [a, b] is divided in small pieces. measure of intervals is its length.
Lebesgue’s approach is to divide the range of the function in small pieces and
then measure the “size” of the pre-image. Hence admissible are functions whose Definition A.4. Let T be a set with a ring Σ of sets. A pre-measure µ on
pre-images of intervals can be measured in some sense. (T, Σ) is a function

µ : Σ → [0, ∞]
A.1 A reminder on measure theory such that µ(∅) = 0 and
Definition A.1. Let T be a set and Σ ⊂ PT a family of subsets of T . [ [ X
(An )n∈N ⊆ Σ, pairwise disjoint and An ∈ Σ =⇒ µ( An ) = µ(An ).
n∈N n∈N n∈N
(i) Σ ⊂ PT is called a ring if for all A, B ∈ Σ also A ∪ B and A \ B belong to
Σ.
S Note that a pre-measure is monotonic: if A, B ∈ Σ and A ⊆ B, then µ(A) ≤
(ii) Σ ⊂ PT is called a σ-ring if it is a ring and n∈N An ∈ Σ for all (An )n∈N ⊆ µ(B).
Σ.
(iii) Σ ⊂ PT is called an algebra if Σ is a ring and T ∈ Σ, that is Example A.5.R Let A be the set of all finite unions of finite intervals and
define µ(A) := R χA dx where χA is the characteristic function of A. Then µ
(a) ∅ ∈ Σ, is a pre-measure on A.
(b) A ∈ Σ =⇒ T \ A ∈ Σ, Proof. Obviously A is a ring, for every A ∈ A the characteristic function χA is
(c) A, B ∈ Σ =⇒ A ∪ B ∈ Σ. Riemann
S integrable and µ(∅) = 0. S Now let (An )Pn∈N ⊆ A be pairwise disjoint
n n
S with n∈N An ∈ A. Obviously, µ( n=1 An ) = n=1 µ(An ) for every n ∈ N.
(iv) Σ ⊂ PT is called a σ-algebra if it is a algebra and n∈N An ∈ Σ for all For n ∈ N define Bn := A \ (A1 ∪ · · · ∪ An ). Obviously, Bn ∈ A and
(An )n∈N ⊆ Σ.
n
[
Note that for A, B ∈ Σ also A ∩ B = A \ (T \ B) ∈ Σ if Σ is an algebra. µ(A) = µ( Ak ) + µ(Bn ).
k=1

Remark. Σ is indeed a ring in the algebraic sense if one sets A + B := (A ∪ S∞


To prove that µ(A) = µ( k=1 Ak ) it suffices to show that lim µ(Bn ) = 0. Fix
B) \ (A ∩ B) and A · B := A ∩ B. n→∞
ε > 0. Since B ∈ A there exists an compact set Cn ⊆ Bn with µ(Bn \ Cn ) <
2−n ε. Let Dn := C1 ∩ · · · ∩ Cn . Then all Dn are compact and Dn ⊆ Cn ⊆ Bn .
Definition A.2. Let T be a set with a σ-algebra Σ. A measure on Σ is a
By construction, B1 ⊇ B2 ⊇ B3 . . . , hence
function µ : Σ → [0, ∞] such that
n
[ n
[
(i) µ(∅) = 0, µ(Bn \ Dn ) = µ(Bn \ (C1 ∩ · · · ∩ Cn ) = µ( (Bn \ Ck )) ≤ µ( (Bk \ Ck ))
S∞ P∞ k=1 k=1
(ii) (An )n∈N ⊆ A with pairwise disjoint An =⇒ µ( n=1 An ) = n=1 µ(An ). n n
X X
≤ µ(Bk \ Ck ) < 2−k ε = ε.
Obviously, the intersection of rings is again a ring and PT is a ring. Hence, given
k=1 k=1
a family U of subsets of T , there exists a smallest ring containing U, namely the T∞ T∞
intersection of all rings that contain U. This ring is called the ring generated by On the other hand, n=1 Dn ⊆ n=1 Bn = ∅. Since all Dn are compact and
U. Analogously the σ-ring, the algebra and the σ-algebra generated by U are D1 ⊇ D2 ⊇ . . . , there exists an K ∈ N such that Dn = ∅, n ≥ K. Hence
obtained. µ(Bn ) = µ(Bn \ Dn ) < ε for all n ≥ k.
132 A.1. A reminder on measure theory Chapter A. Lp spaces 133

P∞
In order to measure all Borel sets, we have to show that the pre-measure of Note that (Bnj )n,j∈N is countable, hence we have proved µ∗ (A) ≤ n=1 µ∗ (An ).

Example A.5 can be extended to the Borel sets. Now let A ∈ A. Clearly, µ(A) S ≤ µ (A) holds. Now fix ε > 0 and choose
(An )n∈N ⊆ A such that A ⊆ n∈N An and
Theorem A.6 (Hahn). Let T be a set, A a ring on T and µ̃ a pre-measure

X
on A. Let Σ(A) be the σ-algebra generated by A. Then
µ(An ) ≤ µ∗ (A) + ε. n ∈ N.
nX
∞ ∞
[ o n=1
µ(A) := inf µ̃(An ) : (An )n∈N ⊆ Σ, A ⊆ An . S
n=1 n=1 Since A = n∈N (An ∩ A) and µ is a pre-measure on A, it follows that
S∞
If µ̃ is σ-finite, i. e., if there exist An ∈ Σ with µ(An ) < ∞ and T = n=1 An ,

X ∞
X
then the extension µ is unique. µ(A) ≤ µ(An ∩ A) ≤ µ(An ) ≤ µ∗ (A) + ε. n ∈ N.
n=1 n=1
For the proof, we first show that µ̃ can be extended to an outer measure µ∗ on
PT . Then, by the lemma of Carathéodory, the restriction of the outer measure Since this is true for all ε > 0, it follows that µ(A) ≤ µ∗ (A).
to the set of the µ∗ -measurable sets is a measure.
Lemma A.9 (Carathéodory). Let µ∗ be an outer measure on PT . Then the
Definition A.7. Let A be a σ-algebra on T . An outer measure on A is a set M of all µ∗ -measurable sets is a σ-algebra and µ∗ is a measure on M.
function µ∗ : A → [0, ∞] such that
Proof. . . . . . .
(i) µ∗ (∅) = 0,
(ii) A, B ∈ A with A ⊆ B =⇒ µ∗ (A) ≤ µ∗ (B). Proof of Theorem A.6. It suffices to show that the set of the µ∗ -measurable sets
S∞ P∞ contains A . . .
(iii) (An )n∈N ⊆ A =⇒ µ∗ ( n=1 An ) ≤ n=1 µ∗ (An ).
A set A ∈ A is called µ∗ -measurable if Hahn’s theorem gives the desired measure on the Borel sets.

µ∗ (Z) = µ∗ (Z ∩ A) + µ∗ (Z \ A), Z ∈ A. Definition A.10 (Lebesgue completion). Let (T, Σ, µ) be a measure space.
A ⊆ T is a zero set if there exists a B ⊆ T with µ(B) = 0 and A ⊆ B (note
Lemma A.8. With the assumptions of Hahn’s theorem (Theorem A.6) that A does not necessarily belong to Σ). The σ-algebra generated by Σ and
the zero sets is called the Lebesgue completion.
nX
∞ ∞
[ o
µ∗ (A) := inf µ̃(An ) : (An )n∈N ⊆ A, A ⊆ An
n=1 n=1
The measure on the completion of the Borel sets in R is the Lebesgue measure,
usually denoted by λ. .
defines an outer measure on PT . In addition, µ(A) = µ∗ (A) for A ∈ A.

Proof. Properties (i) and (ii) of an outer measure are clear. Now let (An )n∈N ⊆ A.2 Integration
PT and ε > 0. Then there exists a family (Bnj )n,j∈N ⊆ A such that An ⊆
S∞ j
j=1 Bn and In the following, I is always an interval in R.

X ε Definition A.11. A function f : I → R is called measurable if for every (a, b) ⊆
µ(Bnj ) ≤ µ∗ (An ) + , n ∈ N.
2n R its preimage f −1 ((a, b)) is a Borel set.
j=1
S S More generally, let (T, ΣT , µT ) and (S, ΣS , µS ) be measure spaces. A function
By construction A := n∈N An ⊆ n,j∈N Bnj and
f : T → S is called measurable if for every U ∈ ΣS also f −1 (U ) ∈ ΣT .

X X
µ∗ (A) ≤ µ(Bnj ) ≤ µ∗ (An ) + ε. Example A.12. Let E be a Borel set. Then the characteristic function χE is
n,j=1 n=1 measurable.
134 A.2. Integration Chapter A. Lp spaces 135

Definition A.13. Let (T, Σ, µ) be a measure space. A function f : T → C is Definition A.16. Let I be an interval in R.
called a simple function if there are Ek ∈ Σ and αk ∈ C such that R
(i) A function f : I → [0, ∞] is called (Lebesgue) integrable if I f dλ < ∞.
n
X
f= αk χEk . (ii) A function f : I → R is called (Lebesgue) integrable if f + := max{f, 0}
k=1
and f − := max{−f, 0} are integrable. In this case
Z Z Z
It is easy to see that simple functions are measurable. Note, however, that the f dλ := f + dλ − f − dλ.
I I I
sum representation of a simple function is not unique.
(iii) A function f : I → C is called (Lebesgue) integrable if Re(f ) and Im(f )
The next theorem lists important properties of measurable functions.
are integrable. In this case
Z Z Z
Theorem A.14. Let I be an interval in R and fn , f, g : I → R be functions. f dλ := Re(f ) dλ + i Im(f ) dλ.
I I I
(i) If f and g are measurable, then so are f +g, f g, fg (if it exists), max{f, g}
and min{f, g}. The Lebesgue integral has the following properties.
(ii) Every continuous function is measurable.
(iii) If all fn are measurable and f is their pointwise limit (i. e. f (t) = lim fn (t) = Lemma A.17. (i) If f, g are Lebesgue integrable and α ∈ K, then
n→∞ Z Z Z
f (t), t ∈ I), then f is measurable. (αf + g) dλ = α f dλ + g dλ.
(iv) If f is measurable, then there exists a sequence (ϕn )n∈N of simple functions I I I
that converges pointwise to f . If in addition f ≥ 0, then the sequence can (ii) If f is Lebesgue integrable then
be chosen such that ϕn (t) ր f (t), t ∈ I. Z Z
(v) If f is measurable and bounded, then there exists a sequence (ϕn )n∈N of
f dλ ≤ |f | dλ.
simple functions that converges uniformly to f . I I

For Lebesgue integrals much stronger convergence theorems hold than for the
The theorem says that the set of the measurable functions are a vector space
Riemann integral. The most important convergence theorems are the following.
and that it is stable under taking pointwise limits.
Next we introduce the integral for positive functions. Theorem A.18 (Monotone convergence theorem). Let (fn )n∈N be a se-
quence of measurable functions fn : I → [0, ∞] with 0 ≤ f1 ≤ f2 ≤ · · · . Then
Definition A.15. Let (T, Σ, µ) be a measure space.
f : I → [0, ∞], f (t) := lim fn (t)
P n→∞
(i) Let f = nk=1 αk χEk with Ek ∈ Σ and αk ∈ [0, ∞] a simple function. We
define its integral as is measurable and
Z Z
Z n
X f dλ = lim fn dλ.
I n→∞ I
f dµ = αk µ(Ek ).
T k=1
The monotone convergence theorem is also called Beppo Levi theorem.
(ii) Let f : T → [0, ∞] be a measurable function. Choose a sequence (ϕn )n∈N A sequence (fn )n∈N converges to f λ-a.e. if the set, where the sequence does
of simple functions with ϕ1 ≤ ϕ2 ≤ . . . that converges pointwise to f . We not converge to f , has measure zero.
define the integral of f by
Z Z Theorem A.19 (Dominated convergence theorem). Let (fn )n∈N be a se-
quence of measurable functions and assume that there exists a measurable func-
f dµ = lim ϕn dµ.
T n→∞ T tion f such that f (t) = lim fn (t) λ-a.e. If there exists an integrable function
n→∞
g with |fn | ≤ g λ-a.e., then f is integrable and
Of course, it must be proved that the integral in (i) does not depend on the sum Z Z
representation of the simple function, and that the limit in (ii) does not depend f dλ = lim fn dλ.
on the chosen sequence of simple functions. I n→∞ I
136 A.3. Lp spaces Chapter A. Lp spaces 137

The dominated convergence theorem is also called Lebesgue’s convergence the- For the proof of the Minkowski inequality, Hölder’s inequality is used.
orem.
1 1
Theorem A.24. Let 1 ≤ p ≤ ∞ and q the conjugated exponent, i. e., p+q = 1.
Theorem A.20 (Fatou’s lemma). Let (fn )n∈N be a sequence of integrable If f ∈ Lp (Ω) and g ∈ Lq (Ω), then f g ∈ L1 (Ω) and
functions and assume that there exists a measurable
R function f such that f (t) =
lim fn (t) λ-a.e. If there exists a C such that I fn dλ ≤ C for all n ∈ N, then kf gk1 ≤ kf kp kgkq .
n→∞
f is integrable and
Z Z Note that Lp (Ω) is only a seminormed space, because there are non-zero func-
f dλ ≤ lim inf fn dλ. tions f with kf kp = 0.
I n→∞ I
Theorem A.25. (Lp (Ω), k · kp ) is complete.
A.3 Lp spaces
Definition A.26. Let Np (Ω) := {f ∈ Lp : kf kp = 0}. Then Lp := Lp (Ω)/N (Ω)
In the following, Ω is always an open subset of Rn . is a complete normed space.

Definition A.21. For 1 ≤ p < ∞ we define Usually an equivalence class [f ] ∈ Lp (Ω) is simply denoted by f .
n Z o
Lp (Ω) := f : Ω → K : f measurable, |f |p dλ < ∞ , Often one is interested in dense subspaces of Lp (Ω).

Z  p1 Theorem A.27. Let 1 ≤ p < ∞ and Ω ∈ Rn open. Then the test functions
kf kp := |f |p , f ∈ Lp (Ω).

C0∞ (Ω) := D(Ω) := {ϕ ∈ C ∞ (Ω) : supp(ϕ) ⊂ Ω is compact}.
Definition A.22. For a measurable function f : Ω → K we define the essential
supremum form a dense subset of Lp (Ω).

ess sup f := inf{C ∈ R : f (t) ≤ C for λ-almost all t},


ess inf f := − ess sup(−f )
and
L∞ (Ω) := {f : Ω → K : f measurable, ess sup |f | < ∞},
kf k∞ := ess sup |f |, f ∈ L∞ (Ω).

It is easy to see that L∞ (Ω) is a vector space. For 1 ≤ p < ∞ this follows from
Z Z Z
|f + g|p dλ ≤ (|f | + |g|)p dλ ≤ (2 max{|f |, |g|})p dλ
Ω Ω Ω
Z Z
≤ 2p max{|f |p , |g|p } dλ ≤ 2p |f |p + |g|p dλ
Ω Ω
= 2p (kf kpp + kgkpp ) < ∞.
That λf ∈ Lp for λ ∈ K and f ∈ Lp is clear.
That k · kp is a seminorm on Lp follows from the Minkowski inequality:

Theorem A.23 (Minkowski inequality). Let 1 ≤ p ≤ ∞ and f, g ∈ Lp (Ω).


Then
kf + gkp ≤ kf kp + kgkp .
Chapter B. Exercises 139

Appendix B

Exercises

Exercises for Chapter 1


1. Banach’s fixed point theorem. Let M be a metric space. A map f :
M → M is called a contraction if there exists a γ < 1 such that

d(f (x), f (y)) ≤ γ d(x, y), x, y ∈ M.

Show that every contraction f on a complete normed space M has exactly


one fixed point, that is, there exists exactly one x0 ∈ M such that f (x0 ) =
x0 .

2. Let X be a normed space. Then the following is equivalent:

(i)X is complete.
(ii)Every absolutely convergent series in X converges in X.

3. Let X be a normed space. Show:

(a) Every finite-dimensional subspace of X is closed.


(b) If V is a finite-dimensional subspace of X and W is a closed subspace
of X, then

V + W := {v + w : v ∈ V, w ∈ W }

is a closed subspace of X.

4. Let T be a set and ℓ∞ (T ) be the space of all functions x : T → K with

kxk∞ := sup{|x(t)| : t ∈ T } < ∞.

Show that (ℓ∞ (T ), k · k∞ ) is a Banach space.


140 Chapter B. Exercises 141

5. Let the sequence spaces d, c0 , c be defined as in Example 1.15. 2. En X = ℓ2 (N) considere el subespacio
(a) Show that (c0 , k · k∞ ) and (c, k · k∞ ) are Banach spaces. U = {(xn )n∈N : xn = 0 excepto para un número finito de ı́ndices n}.
(b) Show that (d, k · k∞ ) is a normed space, but that it is not complete.
Sea V el complemento algebraico de U en X, i. e., U es un subespacio tal
6. Sea X un espacio normado con dim X ≥ 1 y S, T operadores lineales en X que U + V = X y U ∩ V = {0}. Muestre que
tales que ST − T S = id. Muestre que al menos uno de estos operadores no

X
es acotado. Ayuda: Muestre que ST n+1 − T n+1 S = (n + 1)T n .
ϕ : X → K, ϕ(x) = un para x = u + v con u ∈ U, v ∈ V.
n=0
7. Sean X y Y espacios normados con X de dimensión finita. Muestre que
toda función lineal T : X → Y es acotada. es un funcional lineal bien definido y no acotado.
8. (a) Sea X = C([a, b]) con la norma k · k∞ . Muestre que
3. (a) Sea c ⊆ ℓ∞ el conjunto de las sucesiones convergentes. Muestre que el
Z b funcional
T : X → C, Tx = x(t) dt
a ϕ0 : c → K, x = (xn )n∈N 7→ lim xn
n→∞
es un operador lineal y acotado. ¿Cuál es su norma?
(b) Ahora considere X con la norma es continuo y calcule su norma.

Z b 1/p (b) Sea ℓ∞ (N, R) el conjunto de todas las sucesiones acotadas en R con la
kxkp := |x(t)|p dt , x ∈ X, norma del supremo. Muestre que existe ϕ ∈ (ℓ∞ (N, R))′ tal que
a
lim inf xn ≤ ϕ(x) ≤ lim sup xn , x = (xn )n∈N ∈ ℓ∞ .
para 1 ≤ p < ∞. ¿Sigue siendo T acotado? Si es ası́, calcule su norma. n→∞ n→∞

9. Sea 1 ≤ p < ∞. Para z = (zn )n∈N ∈ ℓ∞ sea T : ℓp → ℓp definido por 4. Sea X un espacio normado, f : X → K un funcional lineal no nulo y K =
(T x)n = xn zn para x = (xn )n∈N ∈ ℓp . Muestre que T ∈ L(ℓp ) y calcule kT k. ker f

(a) Muestre que dim(X/K) = 1.


Exercises for Chapter 2
(b) Muestre que f es continuo si y solo si ker f es cerrado.
1. Demuestre el teorema de Hahn-Banach para espacios vectoriales complejos.
5. Un isomorfismo entre espacios normados X y Y es un homeomorfismo lineal.
Pruebe las siguientes afirmaciones.
Sugerencia: Para un espacio vectorial sobre los complejos X muestre que:
(a) Si T : X → Y es un isomorfismo [isométrico] entre los espacios norma-
(a) Sea ϕ : X → R un funcional R-lineal, entonces
dos X y Y , entonces T ′ : Y ′ → X ′ es un isomorfismo [isométrico]. Si
Vϕ : X → C, Vϕ (x) := ϕ(x) − iϕ(ix), X y Y son espacios de Banach, el converso también vale.
(b) Si un espacio normado Y es isomorfo a un espacio de Banach reflexivo
es un funcional C-lineal sobre X con Re Vϕ = ϕ. X, entonces Y es un espacio de Banach reflexivo.
(b) Sea λ : X → C un funcional C-lineal con Re λ = ϕ, entonces Vϕ = λ.
(c) Sea p un funcional sublineal sobre X y ϕ, Vϕ definido como en el punto 6. Sea X un espacio normado separable y (x′n )n∈N una sucesión acotada en X ′ .
anterior, entonces Entonces existe una subsucesión (x′nk )k∈N y x′0 ∈ X ′ tal que

|ϕ(x)| ≤ p(x) ⇐⇒ |Vϕ (x)| ≤ p(x), x ∈ X. lim x′nk (x) = x′0 (x), x ∈ X.
k→∞

(d) kϕk = kVϕ k. Es cierto esto sin la hipótesis de que X sea separable?
142 Chapter B. Exercises 143

Z b
7. Sea X un espacio normado y M un subespacio de X. Sea
(a) Qn (f ) → f (t) dt, n → ∞, para todo f ∈ C[a, b].
a
L = {f ∈ X ′ | f (x) = 0 para todo x ∈ M }. Z b

Muestre que L es un subespacio cerrado de X ′ y que M ′ es isométricamente (b) Qn (p) → p(t) dt, n → ∞, para todo polinomio p : [a, b] → K y
isomorfo a X ′ /L. Pn a (n)
sup k=1 |αk | < ∞.
n∈N
8. Sea X un espacio compacto, CR (X) el conjunto de funciones continuas real-
evaluadas sobre X y Y ⊂ X un subconjunto cerrado.
(a) Considere el mapa ρ : CR (X) → CR (Y ) definido por ρ(f ) = f |Y . Sean X, Y, Z espacios de Banach y T : X ⊇ D(T ) → Y un operador lin-
Muestre que I := ker(ρ) es un subespacio cerrado de CR (X). eal.
(b) Sea ρe : CR (X)/I → CR (Y ) el mapa inducido en el espacio cociente.
Pruebe que ρe es una isometrı́a. (a) Sea S : X ⊇ D(S) → Y un operador lineal. Entonces la suma de
operadores S + T se define como
(c) Demuestre que rg(ρ) es completo.
(d) Use el teorma de Stone-Weierstraß para concluir el teorema de Tietze: D(S + T ) := D(S) ∩ D(T ), (S + T )x := Sx + T x.
Sea X un espacio compacto de Hausdorff y Y ⊆ X un subconjunto
cerrado. Entonces cada función continua f : Y → R tiene una extensión (b) Sea R : Y ⊇ D(R) → Z un operador lineal. Entonces el producto de
continua fe : X → R con kfekC(X) = kf kC(Y ) . operadores o composición RT se define como

9. Muestre que en l1 la convergencia débil y la convergencia en norma coinciden. D(RT ) := {x ∈ D(T ) : T x ∈ D(R)}, (RT )x := R(T x).

5. Sean X, Y, Z espacios de Banach, R ∈ L(X, Y ), T : X ⊇ D(T ) → Y ,


Exercises for Chapter 3 S : Y ⊇ D(S) → Z operadores lineales cerrados. Muestre que:
1. (a) Todo espacio métrico completo con infinitos elementos y ningún punto (a) R + T es un operador lineal cerrado.
aislado es no enumerable.
(b) SR es cerrado.
(b) Toda base algebraica de un espacio de Banach infinito dimensional es
no enumerable. (c) Si S es contı́nuamente invertible (i. e., S −1 : rg(S) → Y existe y es
contı́nuo), entonces ST es cerrado.
2. (a) Sea X un espacio de Banach, Y un espacio normado y (Tn )n∈N ⊆
L(X, Y ). Suponga que para todo x ∈ X el lı́mite T x := lim Tn x existe. Muestre además que estas afirmaciones siguen siendo válidas cambiando
Entonces T ∈ L(X, Y ).
n∈N “cerrado” por “clausurable”
(b) Sean X, Y espacios de Banach, Y reflexivo, y (Tn )n∈N ⊆ L(X, Y ) tal 6. Sea X = ℓ2 (N) y
que (ϕ(Tn x))n∈N converge para todo x ∈ X y ϕ ∈ Y ′ . Entonces existe
w
un T ∈ L(X, Y ) tal que Tn −
→ T. T : X ⊇ D(T ) → X, T x = (nxn )n∈N para x = (xn )n∈N .

3. Muestre que la hipótesis de completitud en el principio de acotación uniforme Diga si T es cerrado con:
es necesaria.
(a) D(T ) = {x = (xn )n∈N ∈ ℓ2 (N) : (nxn )n∈N ∈ ℓ2 (N)},
(n) (n) (n)
4. Sea [a, b] ⊆ R, n ∈ N y tome a ≤ t1 < · · · < tn ≤ b y αk ∈ K, (b) D(T ) = d = {x = (xn )n∈N ∈ ℓ2 (N) : xn 6= 0 para solo finitos n}.
k = 1, . . . , n. Para f ∈ C([a, b]) se define
n 7. Sea X un espacio de Banach, n ∈ N y T un operador lineal densamente
X
Qn (f ) :=
(n)
αk
(n)
f (tk ). definido de X en K n . Muestre que T es cerrado si y solo si T ∈ L(X, Kn ).
k=1
8. Sean X y Y espacios normados y T : X ⊇ D(T ) → Y un operador lineal
Muestre que los siguientes enunciados son equivalentes: cerrado.
144 Chapter B. Exercises 145

(a) Sea K ⊂ X compacto. Muestre que T (K) es cerrado en Y . Exercises for Chapter 4
(b) Muestre que si F es un compacto en Y entonces T −1 (F ) es cerrado en
1. Sea X un espacio pre-Hilbert, U ⊆ H un subespacio denso y x0 ∈ X tal que
X.
hx0 , ui = 0 para todo u ∈ U . Muestre que x0 = 0.
(c) ¿Si A es cerrado en X, es cierto que T (A) es cerrado?
2. Sea w ∈ C([0, 1], R). Para x, y ∈ C([0, 1]) se define
9. Sea X un espacio normado. Una sucesión (xn )n∈N ⊆ X es una sucesión débil Z 1
de Cauchy si para todo ϕ ∈ X ′ la sucesión (ϕ(xn ))n∈N es una sucesión de
hx , yiw := x(t)y(t)w(t) dt.
Cauchy en K. 0

(a) Sea x = (xn )n∈N una sucesión acotada en X. Muestre que x es una Halle una condición necesaria y suficiente spobre w para que h· , ·iw sea
sucesión débil de Cauchy si y solo si existe un subconjunto denso U ′ de un producto interno. Bajo qué condición la norma inducida por h· , ·iw es
X ′ tal que (ϕ(xn ))n∈N es una sucesión de Cauchy para todo ϕ ∈ U ′ . equivalente a la norma usual de L2 ?

(b) Toda sucesión débil de Cauchy en X es acotada. 3. Let H be Hilbert space, (xn )n∈N ⊆ H and x0 ∈ H. Then the following is
equivalent:

10. Sea X un espacio de Banach, (xn )n∈N ⊆ X, (ϕn )n∈N ⊆ X ′ , y x0 ∈ X, (a) xn → x0 .


k·k w
′ w∗
ϕ0 ∈ X tal que xn −−→ x0 y ϕn −−→ ϕ0 . Muestre que lim ϕn (xn ) = ϕ0 (x0 ). (b) kxn k → x0 and xn −
→ x0 .
n→∞
4. Ejemplo de una proyección no acotada. Sea H = l2 y ei el vector usual
eji = δij . Defina
11. Sea X un espacio normado. L1 := span{e2n+1 : n ∈ N0 }
(a) Muestre que (X, k · k)′ = (X, σ(X, X ′ ))′ . Es decir: un funcional lineal y
 
ϕ : X → K es continua con respecto a la topologı́a inducida por k · k si 1 1 1
L2 := span e1 + e2 , e3 + 2 e4 , e5 + 3 e6 , . . . .
y sólo si es continua con respecto a la topologı́a débil. 2 2 2
w
(b) Sean (xn )n∈N ⊆ X, x0 ∈ X y (ϕn )n∈N ⊆ X ′ , ϕ0 ∈ X ′ tal que xn −
→ x0
w∗ (a) Muestre que L1 ∩ L2 = {0}.
y ϕn −−→ ϕ0 . Muestre
(b) Muestre que L1 ⊕ L2 = H.
kx0 k ≤ lim inf kxn k, kϕ0 k ≤ lim inf kϕn k. (c) Muestre que L1 ⊕ L2 6= H.
n→∞ n→∞
(d) Defina el operador P0 : L1 ⊕ L2 → L1 ⊕ L2 , P0 (x + y) = x. Muestre
(c) Sean S = {x ∈ X : kxk = 1} la esfera unitaria y K = {x ∈ X : kxk = que P0 es una proyección no acotada.
1} la bola unitaria cerrada en X. ¿Siempre son débilmente cerradas
(prueba o contraejemplo)? 5. Para λ ∈ R defina fλ : R → C, fλ (s) = eiλs y sea X = span{fλ : λ ∈ R}.
Muestre que
Z T
12. Para n ∈ N sea en = (0, . . . , 1, 0, . . . ) la sucesión que tiene 1 en la posición 1
hf , gi := lim f (s)g(s) ds
n y 0 en el resto. T →∞ 2T −T

(a) Muestre que (en )n∈N no es convergente débilmente en ℓ1 . define un producto interior en X. Muestre que la completación de X no es
separable. (kfλ − fλ′ k =?)
(b) Muestre que (en )n∈N es w∗ convergente en ℓ1 .
Los elementos en la completación de X se llaman funciones casi periódicas.
13. Sea X un espacio vectorial y M ⊆ X un subconjunto convexo, balanceado
y absorbente. Muestre que el funcional de Minkowski pM es una seminorma 6. ¿Existe algún producto interno h· , ·i en C[0, 1] tal que hx, xi = kxk2∞ para
en X. todo x ∈ C[0, 1] ?
146 Chapter B. Exercises 147

7. Sea X un espacio pre-Hilbert. Muestre los siguientes resultados W m (Ω) es un producto interior con
X
(a) Sean x, y ∈ X con x ⊥ y, entonces hf , giW m := hD(α) f , D(α) gi2 .
|α|≤m
kx + yk2 = kxk2 + kyk2 .

¿El converso es cierto en general? ¿Hay algún caso para el que se tenga? Además, definimos los espacios
(b) Si x 6= 0, y 6= 0 y x ⊥ y muestre que el conjunto {x, y} es linealmente H m (Ω) := C m (Ω) ∩ W m (Ω) and H0m (Ω) := D(Ω)
independiente.
¿Como se puede generalizar este resultado? donde la clausura es tomada con respecto a la norma inducida por h· , ·iW m .
(c) x ⊥ y, si y solo si kx + αyk ≥ kxk para todo escalar α.
p B : H × H → K sesquilineal. En H × H
12. Sea H un espacio de Hilbert and
considere la norma k(x, y)k := kxk2 + kyk2 .
8. Let H be a Hilbert space, Y ⊆ H a subspace and ϕ0 ∈ Y ′ . Show that there
exists exactly one extension ϕ ∈ H ′ of ϕ0 with kϕ0 k = kϕk. (a) Muestre que las siguientes son equivalentes:

9. Sea X un espacio pre-Hilbert y U ⊆ X un subespacio. Muestre que (i) B es contı́nua.


(ii) B is parcialmente contı́nua, es decir, para cada x0 fijo, y 7→ B(x0 , y)
⊥⊥
(a) U 6= U . ¿Se tiene alguna contenencia? es contı́nua para cada y0 fijo, x 7→ B(x, y0 ) es contı́nua.
(b) U ⊕ U ⊥ 6= X (iii) B es acotado, es decir, existe M ∈ R tal que kB(x, y)k ≤ M kxkkyk
para todo x, y ∈ H.
10. Sea 1 ≤ p ≤ ∞. Para f ∈ Lp (R) y s ∈ R defina Ts : Lp (R) → Lp (R) como (b) Si B es contı́nuo, entonces existe T ∈ L(H) tal que
(Ts f )(t) := f (t − s). Claramente los Ts son isometrias lineales.
s B(x, y) = hT x , yi, x, y ∈ H.
(a) Sea 1 ≤ p < ∞. Muestre que Ts −
→ id para s → 0. Los Ts convergen
en norma? (c) Si además existe m > 0 tal que B(x, x) ≥ mkxk2 , x ∈ H, entonces T
(b) Los Ts convergen en norma o convergen fuertemente en el caso p = ∞? es invertible y kT −1 k ≤ m−1 .

11. Muestre que W m (Ω), H m (Ω) y H0m (Ω) son espacios de Hilbert. 13. Sea H un espacio de Hilbert. Muestre que para toda sucesión (xn )n ⊆ H
acotada, existe una subsucesión (xnk )k tal que la sucesión (ym )m donde,
Para el problema 4.10: Para Ω ⊆ R definimos el conjunto de funciones de
prueba m
1 X
ym = xnk ,

D(Ω) := {ϕ ∈ C (Ω) : supp(ϕ) ⊆ Ω es compacto}. m
k=1

converge.

Para un multi-ı́ndice α = (α1 , . . . , αn ) ∈ Nn se define |α| = α1 + · · · + αn y 14. Sea X un espacio normado, (xn )n∈N ⊆ X y x ∈ X. Las siquientes son
Dα ϕ = ∂1α1 . . . ∂nαn ϕ si la derivada existe. equivalentes:
P
Sea f ∈ L2 (Ω). Una función g ∈ L2 (Ω) se llama la derivada débil α-ésima (a) n∈N xn converge incondicionalmente a x.
de f si
(b) Para todo ε > 0 existe un conjunto finito A ⊆ N tal que para todo
hg , ϕi = (−)|α| hf , Dα ϕi, ϕ ∈ D(Ω). conjunto finito B con A ⊆ B ⊆ N
X

Note que la derivada débil es única si existe; se denota por D(α) f . xb − x < ε.
b∈B
Para m ∈ N definimos el espacio de Sobolev
15. Sea H un espacio de Hilbert. Si P : H → H es un operador lineal, las
W m (Ω) := {f ∈ L2 (Ω) : D(α) f ∈ L2 (Ω), |α| ≤ m}. siquientes son equivalentes:
148 Chapter B. Exercises 149

(a) P es una proyección ortogonal. (a) (hj )j∈N0 is a orthonormal system in L2 [0, 1] and (ψn,k )n,k∈Z is a or-
(b) P 2 = P y hP x , yi = hx , P yi. thonormal system in L2 (R).
P2k −1
(b) T : L2 [0, 1] → L2 [0, 1], T f = j=0 hf , hj ihj is a orthonormal projec-
16. Sea H un espacio de Hilbert, V, W ⊆ H subespacios cerrados y PV , PW sus
tion on the subspace
correspondientes proyecciones ortogonales.
(a) Muestre que U = {f ∈ L2 [0, 1] : f const. in intervals [r2−k , (r + 1)2−k ) with r ∈ N0 }.
P∞
V ⊆W ⇐⇒ PV = PV PW = PW PV . (c) For f ∈ C[0, 1], the series j=0 hf , hj ihj converges uniformely to f .
(b) Muestre que las siguientes afirmaciones son equivalentes: (d) (hj )j∈N0 is an orthonormal basis of L2 [0, 1].
(i) PV PW = 0. (e) (ψk,n )k,n∈Z is an orthonormal basis of L2 (R).
(ii) V ⊥ W .
(iii) PV + PW es una proyección ortogonal. 21. Sea H un espacio de Hilbert, V, W ⊆ H subespacios cerrados y PV , PW sus
correspondientes proyecciones ortogonales.
Muestre que rg(PV + PW ) = V ⊕ W si alguna de las condiciones anteriores
se tiene. (a) Muestre que

17. Sea H un espacio de Hilbert y P0 , P1 las proyecciones ortogonales sobre V ⊆W ⇐⇒ PV = PV PW = PW PV .


H0 , H1 ⊆ H. Entonces las siguientes afirmaciones son equivalenteas:
(b) Muestre que las siguientes afirmaciones son equivalentes:
(a) H0 ⊆ H1 ,
(i) PV PW = 0.
(b) kP0 xk ≤ kP1 xk, x ∈ H.
(ii) V ⊥ W .
(c) hP0 x , xi ≤ hP1 x , xi, x ∈ H. (iii) PV + PW es una proyección ortogonal.
(d) P0 P1 = P0 .
Muestre que rg(PV + PW ) = V ⊕ W si alguna de las condiciones anteriores
18. Sea H un espacio de Hilbert separable, (xn )n∈N una base ortonormal de H, se tiene.
y, (yn )n∈N una sucesión tal que:
∞ 22. Sea H un espacio de Hilbert y P0 , P1 las proyecciones ortogonales sobre
X
||xn − yn || < 1 H0 , H1 ⊆ H. Entonces las siguientes afirmaciones son equivalenteas:
n=1
(i)H0 ⊆ H1 ,
y z ⊥ yn , para todo n ∈ N, entonces z = 0.
(ii)kP0 xk ≤ kP1 xk, x ∈ H.
19. Sea H un espacio de Hilbert complejo y T : H → H un operador lineal (iii)hP0 x , xi ≤ hP1 x , xi, x ∈ H.
acotado. Muestre que T es normal si y solo si ||T ∗ x|| = ||T x|| para todo (iv)P0 P1 = P0 .
x ∈ H. En este caso, muestre que ||T 2 || = ||T ||2 .
23. Sea H un espacio de Hilbert separable, (xn )n∈N una base ortonormal de H,
20. Haar functions. Let ψ = χ[0, 1/2) − χ[1/2, 1) . For n, k ∈ Z define y, (yn )n∈N una sucesión tal que:
ψn,k : R → R, ψn,k (t) = 2k/2 ψ(2k t − n). ∞
X
||xn − yn || < 1
For k ∈ N0 and n ∈ {0, 1, 2, . . . , 2k − 1} let n=1


h2k +n (t) = ψk,n (t), for t ∈ [0, 1), y z ⊥ yn , para todo n ∈ N, entonces z = 0.
h2k +n : [0, 1] → R,

h2k +n (1) = lim ψk,n (t). 24. Sea H un espacio de Hilbert complejo y T : H → H un operador lineal
t→1−
acotado. Muestre que T es normal si y solo si ||T ∗ x|| = ||T x|| para todo
and h0 (t) = 1, t ∈ [0, 1]. x ∈ H. En este caso, muestre que ||T 2 || = ||T ||2 .
150 Chapter B. Exercises 151

25. Sea H un espacio de Hilbert y (Tn )n∈N una sucesión acotada y monótonamente 2. Sea (λn )n∈N ⊆ C una sucesión acotada, y,
creciente de operadores autoadjuntos. Muestre que la sucesión converge en
el sentido fuerte a un operador autoadjunto. T : ℓ1 → ℓ1 , T ((xn )n∈N ) = (λn xn )n∈N .

26. Sea (Pn )n∈N una sucesión monótona de proyecciones ortogonales en un es- Encuentre σ(T ), σp (T ), σc (T ) y σr (T ). Muestre además que, para todo
pacio de Hilbert H. Muestre que (Pn )n∈N converge en el sentido fuerte a K ⊆ C compacto no vacı́o, existe un operador T ∈ L(ℓ1 ) cuyo espectro es
una proyección ortogonal P y además K.
S
(a) rg P = n∈N rgPn si Pn es creciente. 3. Sea X un espacio de Banach S, T ∈ L(X). Muestre que σ(ST ) \ {0} =
T σ(T S) \ {0}.
(b) rg P = n∈N rgPn si Pn es decreciente.
Hint. Muestre que id −ST es invertible si y solo si id −T S es invertible, en-
27. Sean H1 , H2 y H3 espacios de Hilbert y S(H1 → H2 ) y T (H2 → H3 )
contrando una relación entre (id −T S)−1 y (id −ST )−1 . Suponga ||T || ||S|| <
operadores lineales densamente definidos.
1 y mire si la relación en este caso es válida en general.
(a) Si T ∈ L(H2 , H3 ) entonces T S es densamente definido y (T S)∗ = S ∗ T ∗ .
4. Encuentre el espectro puntual, el espectro continuo y el espectro residual de
los operadores:
(b) Si S es inyectivo y S −1 ∈ L(H2 , H1 ) entonces T S es densamente
definido y (T S)∗ = S ∗ T ∗ . R : ℓ2 (N) → ℓ2 (N), R(x1 , x2 , x3 , . . . ) = (0, x1 , x2 , x3 , . . . ),
(c) Si S es inyectivo y S −1 ∈ L(H2 , H1 ) entonces S ∗ es inyectivo y (R∗ )−1 = L : ℓ2 (N) → ℓ2 (N), L(x1 , x2 , x3 , . . . ) = (x2 , x3 , x4 , . . . ),
(R−1 )∗
T : ℓ∞ (N) → ℓ∞ (N), T (x1 , x2 , x3 , . . . ) = (x2 , x3 , x4 , . . . ).
28. Sean H1 , H2 espacios de Hilbert y U : H1 × H2 → H2 × H1 , U (x, y) =
(−y, x). Entonces
(a) U es unitario.
(b) Si T (H1 → H2 ) es densamente definido,
G(T ∗ ) = [U (G(T ))]⊥ = U (G(T )⊥ ).

(c) T ∗ es cerrado.
(d) Si T es clausurable, T ∗ es densamente definido y T ∗∗ = T .

Exercises for Chapter 5


1. (a) Sea X = C([0, 1]) y a ∈ C([0, 1]). Muestre que
A : X → X, (Ax)(t) = a(t)x(t)
es un operador lineal acotado. Encuentre kAk, σ(A), σp (A), σc (A) y
σr (A).
(b) Sea H = {f ∈ C([0, 1]) : x(0) = 0} y
Z t
S : H → H, (Sx)(t) = x(s) ds.
0

Encuentre σ(S), σp (S), σc (S) y σr (S).


Bibliography

[Den04] Robert Denk. Funktionalanalysis I. Universität Konstanz,


http://www.mathematik.uni-regensburg.de/broecker/index.html,
2004.
[DS88] Nelson Dunford and Jacob T. Schwartz. Linear operators. Part I. Wiley
Classics Library. 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.
[Hal98] Paul R. Halmos. Introduction to Hilbert space and the theory of spectral
multiplicity. AMS Chelsea Publishing, Providence, RI, 1998. Reprint
of the second (1957) edition.
[Heu06] Harro Heuser. Funktionalanalysis. Mathematische Leitfäden. [Math-
ematical Textbooks]. B. G. Teubner, Stuttgart, fourth edition, 2006.
Theorie und Anwendung. [Theory and application].
[Kat95] Tosio Kato. Perturbation theory for linear operators. Classics in Math-
ematics. Springer-Verlag, Berlin, 1995. Reprint of the 1980 edition.
[Les] Jaime Lesmes. Análisis funcional. 2 edición.
[Mac09] Barbara D. MacCluer. Elementary functional analysis, volume 253 of
Graduate Texts in Mathematics. Springer, New York, 2009.
[RS80] Michael Reed and Barry Simon. Methods of modern mathematical
physics. I. Academic Press Inc. [Harcourt Brace Jovanovich Publish-
ers], New York, second edition, 1980. Functional analysis.
[Rud87] Walter Rudin. Real and complex analysis. McGraw-Hill Book Co., New
York, third edition, 1987.
[Rud91] Walter Rudin. Functional analysis. International Series in Pure and
Applied Mathematics. McGraw-Hill Inc., New York, second edition,
1991.
[Wei80] Joachim Weidmann. Linear operators in Hilbert spaces, volume 68
of Graduate Texts in Mathematics. Springer-Verlag, New York, 1980.
Translated from the German by Joseph Szücs.

153
154 Bibliography

[Wer00] Dirk Werner. Funktionalanalysis. Springer-Verlag, Berlin, extended


edition, 2000.
[Yos95] Kōsaku Yosida. Functional analysis. Classics in Mathematics. Springer-
Verlag, Berlin, 1995. Reprint of the sixth (1980) edition.
[Zei95] Eberhard Zeidler. Applied functional analysis, volume 109 of Applied
Mathematical Sciences. Springer-Verlag, New York, 1995. Main prin-
ciples and their applications.

Das könnte Ihnen auch gefallen