Functional Analysis
Functional Analysis
Analysis
Series
M. Winklmeier
Chigüiro Collection
Contents
1 Banach spaces 7
1.1 Metric spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2 Normed spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.3 Hölder and Minkowski inequality . . . . . . . . . . . . . . . . . . 22
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
3
4 CONTENTS
A Lp spaces 129
A.1 A reminder on measure theory . . . . . . . . . . . . . . . . . . . 130
A.2 Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
A.3 Lp spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
B Exercises 139
References 152
CONTENTS 5
Notation
The letter K usually denotes either the real field R or the complex field C. The
positive real numbers are denoted by R+ := (0, ∞).
Chapter 1. Banach spaces 7
Chapter 1
Banach spaces
d:M ×M →R
(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,
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
such that Bε (p) ⊆ U . In particular, the open balls are open and closed balls are
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 .
It is easy to see that the topology generated by d has the Hausdorff property,
that is, for every x 6= y ∈ M there exist neighbourhoods Ux of x and Uy of y
with Ux ∩ Uy = ∅.
Recall that a set N ⊆ M is called dense in M if N = M , where N denotes the
closure of N .
Not every metric space is complete, but every metric space can be completed in
the following sense.
Theorem 1.7. Let (M, d) be a metric space. Then there exists a complete
metric space (M b and an isometry ϕ : M → M
c, d) c such that ϕ(M ) = M
c. M
c is
called completion of M ; it is unique up to isometry.
Proof. Let
x∼y ⇐⇒ d(xn , yn ) → 0, n → ∞
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)
and (ii) of the definition of a metric and transitivity of ∼ is a consequence of
the triangle inequality).
Let Mc := CM / ∼ the set of all equivalence classes. The equivalence class
containing x = (xn )n∈N is denoted by [x]. On M c we define
c×M
db : M c → R, d([x],
b [y]) = lim d(xn , yn ). (1.1)
n→∞
Since (d(xn , yn ))n∈N is a Cauchy sequence in the complete space R, the limit in
(1.1) exists.
10 1.1. Metric spaces
Moreover, for (x̃n )n∈N ∈ [x] and (ỹn )n∈N ∈ [y] it follows that
Hence db is well-defined.
Let
ϕ:M →M
c, ϕ(x) = [(x)n∈N ].
• 0 = d([x],
b [y]) = lim d(xn , yn ) ⇐⇒ x∼y ⇐⇒ [x] = [y].
n→∞
• d([x],
b [y]) = lim d(xn , yn ) = lim d(yn , xn ) = d([y],
b [x]).
n→∞ n→∞
• d([x],
b [y]) = lim d(xn , yn ) ≤ lim d(xn , zn ) + d(zn , yn ) = d([x],
b [z]) +
n→∞ n→∞
d([z],
b [y]).
Step 2: ϕ is an isometry.
Proof. This follows immediately from the definition.
Step 3: ϕ(M ) = M c.
Proof. Let (xn )n∈N ∈ [x] ∈ M c and ε > 0. Then there exists an N ∈ N such that
ε
d(xn , xm ) < 2 , m, n ≥ N . Let z := xN ∈ M . Then
ε
d(ϕ(z),
b [x]) = lim d(xN , xn ) ≤ < ε.
n→∞ 2
b xn , zn ) < 1 ,
d(b n ∈ N.
n
The sequence z is a Cauchy sequence in M because
b n , z) ≤ d(x̂ 1
d(x̂ b n , ϕ(zn )) + d(ϕ(z
b n ), z) < + lim d(zn , zm ) → 0, n → ∞.
n m→∞
Chapter 1. Banach spaces 11
We have shown that ϕ(M ) is a dense subset of the complete metric space (M
c, d)
b
and that ϕ is an isometry.
Finally, we have to show that Mc is unique (up to isometry). Let (N, dN ) be
complete metric space and ψ : M → N an isometry such that ψ(M ) = N . Then
the map
T : ϕ(M ) → ψ(M ), T (ϕ(x)) = ψ(x)
c → N by
can be extended to a surjective isometry T : ϕ(M ) = M
T x = T ( lim xn ) := lim T xn
n→∞ n→∞
Then d1 and d2 are metrics on C([a, b]). (C([a, b]), d1 ) is complete, (C([a, b]), d2 )
is not complete.
Remark. The completion of (C([a, b]), d2 ) is L1 (a, b) (the set of all Lebesgue
integrable functions on (a, b)).
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
on M there exists a countable set A := {xn : n ∈ N} ⊆ M such that A = M .
1
Let J := {(n, m) ∈ N × N : ∃y ∈ N with d(xn , y) < m }. For every (n, m) ∈ J
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
every y ∈ N and k ∈ N there exists a b ∈ B such that d(b, y) < k1 . By definition
1
of A there exists a xn ∈ A such that d(xn , y) < 2k . In particular, (n, 2k) ∈ J.
It follows that d(yn,2k , y) ≤ d(yn,2k , xn ) + d(xn , y) < k1 .
12 1.2. Normed spaces
k·k:X →R
(i) kxk = 0 ⇐⇒ x = 0,
(ii) kαxk = |α| kxk,
(iii) kx + yk ≤ kxk + kyk.
Remarks. • Note that the implication ⇐ in (i) follows from (ii) because
k0k = k2 · 0k = 2k0k.
d(x, y) := kx − yk, x, y ∈ X.
Hence a norm induces a topology on X via the metric and we have the concept
of convergence etc. on a normed space.
(i) X is complete.
(ii) Every absolutely convergent series in X converges in X.
x∼y ⇐⇒ x − y ∈ M.
Chapter 1. Banach spaces 13
Examples 1.15. (i) Finite dimensional normed spaces. Cn and Rn are com-
plete normed spaces with
k · k∞ : Kn → R, kxk∞ = max{|xj | : j = 1, . . . , n}.
The triangle inequality kx+ykp ≤ kxkp +kykp is called the Minkowski inequality
(see Section 1.3).
(ii) Let T be a set and define
`∞ (T ) := {x : T → K bounded map}.
Obviously, `∞ (T ) is a vector space. Let
kxk∞ := sup{|x(t)| : t ∈ T }, x ∈ `∞ ,
be the supremum norm. Then (`∞ (T ), k · k∞ ) is a Banach space.
and
∞
X p1
kxkp := |xn |p , x ∈ `p .
n=1
and
∞ ∞ ∞
X X p X p
|xn + yn |p ≤ 2 max{|xn |, |yn |} = 2p max{|xn |, |yn |}
n=1 n=1 n=1
∞
X
≤ 2p |xn |p + |yn |p = 2p (kxkpp + kykpp ) < ∞.
n=1
Since K is complete, the limit ym := lim xn,m exists. Let y := (ym )m∈N .
n→∞
k·kp
We will show that y ∈ `p and that xn −−−→ y. Let ε > 0 and N ∈ N such
that kxn − xk k < ε for all k, n ≥ N . For every M ∈ N
M
X
|xn,j − xk,j |p ≤ kxn − xk kpp < εp .
j=1
C(T ) := {f : T → K : f is continuous},
B(T ) := {f : T → K : f is bounded},
BC(T ) := C(T ) ∩ B(T ).
kf k∞ := sup{|f (t)| : t ∈ T }.
kf k(1) := kf k∞ + kf 0 k∞ .
Then (C 1 ([a, b]), k · k(1) ) is a Banach space. Note that the right hand side
is finite because by assumption f 0 is continuous.
Proof. Let (xn )n∈N be a Cauchy sequence in (C 1 ([a, b]), k · k(1) ). Then
there exist x, y ∈ C([a, b]) such that xn → x and x0n → y in the supremum
norm. A well-known theorem in analysis implies x0 = y, hence xn → x in
k · k(1) .
Chapter 1. Banach spaces 17
1
1
z + y nk
=
xnk
→ 0, n → ∞.
ank ank
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].
Obvioulsy, U and V are closed subspaces of `1 . Let en be the nth unit vector
1
in `1 . Let m ∈ N. Then e2m−1 ∈ U ⊆ V + U and e2m = (e2m + m e2m−1 ) −
1
m e2m−1 ∈ V + U . Since span{en : n ∈ N} is a dense subset of ` 1 , it follows
that V + U = `1 .
Now we will show that V + U 6= `1 . Let
(
1
x = (xn )n∈N , xn = n2 , n even,
0, n odd.
The theorem above implies in particular, that the topologies generated by equiv-
alent norms coincide. Moreover, the identity map id : (X, k · k1 ) → (X, k · k2 ) is
uniformly continuous for equivalent norms.
and let k · k(1) be as in Example 1.15 (7). It is not hard to see that
The theorem above implies that all norms a a finite-dimensional K-vector space
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
dimensional normed space is compact if and only if it is bounded and closed
(Theorem of Heine-Borel for Kn with the Euclidean metric). In particular, the
unit ball in a finite dimensional space is compact.
This is no longer true in infinite dimensional normed spaces. In fact, the unit
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
shows that in a certain sense quotient spaces can work as a substitute for the
orthogonal complement in inner product spaces (see 4.2).
Proof. If Y = {0} or ε ≥ 1, the assertion is clear. Now assume 0 < ε < 1. Note
1
that in this case 1−ε > 1. Since Y is closed and different from X, the quotient
space X/Y is not trivial. Hence there exists an ξ ∈ X such that k[ξ]k∼ = 1. By
Remark 1.14 there exists y ∈ Y such that
1
1 = k[ξ]k∼ ≤ kξ + yk < .
1−ε
Let x = kξ + yk−1 (ξ + y). Obviously, kxk = 1 and for every z ∈ Y
kx − zk = kξ + yk−1
ξ + y − kξ + ykz
≥ kξ + yk−1 k[ξ]k∼ = kξ + yk−1 > 1 − ε.
| {z }
∈Y
Let X be a vector space and Λ a set. A family (xλ )λ∈Λ ⊆ X is called linearly
independent if every finite subset is linearly independent. A Hamel basis (or an
Chapter 1. Banach spaces 21
span(Y ) = X,
Note that every normed space with a Schauder basis is separable, but not every
separable normed space has a Schauder basis.
22 1.3. Hölder and Minkowski inequality
Proof. Recall that the set A := {(xn )n∈N : xn ∈ {0, 1}} i s not countable.
Obviously, A ⊆ `∞ . Let B be a dense subset of `∞ . 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
follows that B has at least the cardinality of A, that is, there exists no countable
dense subset of `p .
(iii) C[a, b] is separable since by the theorem of Weierstraß the set of polyno-
mials
{[a, b] → R, x 7→ xn : n ∈ N}
Proof. If ab = 0, then inequality (1.5) is clear. Now assume ab > 0. Since the
logarithm is concave and p1 + 1q = 1 is follows that
1 1 q 1 1
ln ap + b ≥ ln(ap ) + ln(bq ) = ln(a) + ln(b) = ln(ab).
p q p q
p
Theorem 1.28 (Hölder’s inequality). Let 1 ≤ p ≤ ∞ and q = p−1 , i. e.,
1 1
+ =1
p q
1
(setting ∞ = 0). If x ∈ `p and y ∈ `q , then z = (xn yn )n∈N ∈ `1 and
Proof. If x = 0 or y = 0 then the inequality (1.6) clearly holds. Also the cases
p = 1 and p = ∞ are clear.
Now assume x, y 6= 0 and 1 < p < ∞. The Young inequality (1.6) with
|xj | |yj |
a= , b=
kxkp kykq
yields
M
X M
X
p
|xj + yj | = |xj + yj | · |xj + yj |p−1
j=1 j=1
M
X M
X
p−1
≤ |xj | |xj + yj | + |yj | |xj + yj |p−1
j=1 j=1
p p
M
X M
1 X z }| { 1 M
X M
p1 X z }| { 1
p q (p−1)q q
≤ |xj |p |xj + yj |(p−1)q + |yj |p |xj + yj |
j=1 j=1 j=1 j=1
M
X 1
p q
≤ kxkp + kxkp |xj + yj | .
j=1
P 1
M p q
Note that j=1 |x j + yj | 6= 0 for M large enough. Hence the above
inequality yields
M
X p1
|xj + yj |p ≤ kxkp + kxkp
j=1
p 1
using p − q =p 1− q = 1. Taking the limit M → ∞ finally proves (1.7).
Chapter 2. Bounded maps; the dual space 25
Chapter 2
(i) T : X → Y is continuous
(ii) lim T xn = T limn→∞ xn for every convergent sequence (xn )n∈N ∈ X
n→∞
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 xk ≤ kT k kxk, x ∈ X.
26 2.1. Bounded linear maps
kT k = sup{kT xk : x ∈ X, kxk = 1}
= sup{kT xk : x ∈ X, kxk ≤ 1}
n kT xk o
= sup : x ∈ X, x 6= 0
kxk
= inf{M ∈ R : ∀ x ∈ X kT xk ≤ M kxk}.
ST : X → Z, ST x := S(T x).
Proof. (i) In Remark 2.4 we have seen that L(X, Y ) is a vector space. From
definition of the operator norm it is clear that kT k = 0 if and only if T = 0
and that kλT k = |λ| kT k for all λ ∈ K. To prove the triangle inequality let
S, T ∈ L(X, Y ) and x ∈ X.
T : X → Y, T x := lim Tn x.
n→∞
Examples 2.7. In the following examples, the linearity of the operator under
consideration is easy to check.
kT k = sup kT
xk∞ 1
kT xn k∞
kxkC 1 : x ∈ C ([0, 1]) \ {0} ≥ sup kxn kC 1 : n ∈ N
= sup 1+1 1 : n ∈ N = 1.
n
1
Proof. As in the example above let xn : [0, 1] → R, xn (t) := n exp(−nt).
It follows that
kT xk∞ kT xn k∞
sup kxk∞ : x ∈ C 1 ([0, 1]) \ {0} ≥ sup :n∈N
kxn k∞
= sup 11 : n ∈ N = ∞
n
Xn
n
X n
X
kT xkY =
αj T ej
≤ |αj | kT ej kY ≤ M |αj | = M kxkX .
Y
j=1 j=1 j=1
Chapter 2. Bounded maps; the dual space 29
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
continuous.
Assume that S is an arbitrary continuous extension of T . For x ∈ X and a
sequence (xn )n∈N ⊆ D which converges to x we find
Proof. The proof is analogous to the proofPfor the convergence of the geometric
m
series. We define the partial sums Sm := n=0 T n , m ∈ N0 . Then
Note that:
30 2.1. Bounded linear maps
P∞
(i) T m → 0 for m → ∞ because m=0 T m converges.
P∞
(ii) Sm → n=0 T n for m → ∞ by assumption.
(iii) For fixed R ∈ L(X) the maps L(X) → L(X), S 7→ RS and S 7→ SR
respectively are continuous.
has solution x ∈ C([0, 1]). If a solution exists, is it unique? Can the norm of
the solution be estimated in terms of y?
Solution. Note that equation (2.4) can be written as an equation in the Banach
space C([0, 1]):
x − Kx = y
where
Z s
K : C([0, 1]) → C([0, 1]), (Kx)(s) := k(s, t)x(t) dt, s ∈ [0, 1].
0
Obviously, K is a well-defined linear operator and for all x ∈ C([0, 1]), s ∈ [0, 1]
Z s Z s
|Kx(s)| = k(s, t)x(t) dt ≤ |k(s, t)| |x(t)| dt ≤ s kkk∞ kxk∞ ,
0 0
Z s Z t Z sZ t
2
k(t, t1 )x(t1 ) dt1 dt ≤ kkk2∞ kxk∞
|K x(s)| = k(s, t) dt1 dt
0 0 0 0
s2
= kkk2∞ kxk∞ .
2
Chapter 2. Bounded maps; the dual space 31
sn
|K n x(s)| ≤ kkkn∞ kxk∞ , s ∈ [0, 1], x ∈ C([0, 1]), n ∈ N,
n!
P∞
which shows that kK n k ≤ kkk ∞
n! . In particular, n=0 K
n
converges so that
id −K is invertible by Theorem 2.10. Hence equation (2.4) has exactly one
solution x ∈ C([0, 1]), given by
∞
X
x= K n y.
n=0
P
∞ P∞ P∞ kkk∞
Moreover, kxk∞ =
n=0 K n y
≤ kK n k kyk∞ ≤ n! kyk∞ =
n=0 n=0
∞
e kkk∞ kyk∞ .
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.
Theorem 2.6 implies immediately:
Proposition 2.13. The dual space of a normed space X with the norm
is a Banach space.
p(x) ≤ M kxk, x ∈ X.
If p satisfies
X → R, x 7→ |ϕ(x)|
is sublinear.
Remark. Observe that p(x) ≥ 0 for every x ∈ X and every sublinear functional
p. Moreover, note that every seminorm is a sublinear functional.
The next fundamental theorem shows that every normed space admits non-
trivial functionals (except when X = {0}).
ϕ0 (y) ≤ p(y), y ∈ Y.
for some c ∈ R. We have to find c such that |ψc (z)| ≤ p(z), z ∈ Z, that is,
By assumption on ϕ0
implying
so that
Proof. (i) For all ϕ ∈ X 0 with kϕk = 1: kxk = kϕk kxk ≥ |ϕ(x)|, hence
kxk ≥ sup{ϕ(x) : ϕ ∈ X 0 , kϕk = 1}. To show that in fact we have equality, we
recall that by Corollary 2.17 there exists a ϕ ∈ X 0 with kϕk = 1 and ϕ(x) = kxk.
Hence the formula in (i) is proved. Note the the supremum is in fact a maximum.
(i) Y = X,
ϕ ∈ X 0.
(ii) ϕ|Y = 0 =⇒ ϕ = 0 ,
X 0 separable =⇒ X separable.
kT xk ≤ kxkq , x ∈ `q . (2.7)
Using pq − p = q we find
N
X N
X
|tn |p = |xn |p(q−1) , N ∈ N.
n=1 n=1
For N large enough, the last factor in the line above is not zero, so, using
1 − p1 = 1q , we obtain
N
X q1
|xn |q ≤ ky 0 k
n=1
Other important examples are given without proof in the following theorems.
Theorem 2.23. Let (Ω, Σ, µ) be a σ-finite measure space. Let 1 ≤ p < ∞ and
q such that p1 + 1q = 1. Then
Z
T : Lq (Ω) → (Lp (Ω))0 , (T f )(g) = f g dµ, f ∈ Lq (Ω), g ∈ Lp (Ω),
Ω
38 2.4. The Banach space adjoint and the bidual
is an isometric isomorphism.
1 1
Let 1 ≤ p < ∞ and q such that p + q = 1. Then
Z
T : M (K) → (C(K))0 , (T µ)(g) = g dµ, µ ∈ M (K), g ∈ C(K),
Ω
is an isometric isomorphism.
(Lp )0 ∼
= Lq , 1 ≤ p < ∞,
0 ∼
(C(K)) = M (K).
T 0 : Y 0 → X 0, (T 0 y 0 )x := y 0 (T x), y 0 ∈ Y 0 , x ∈ X.
T
X Y
x0 =y 0 ◦T y0
(i) The map L(X, Y ) → L(Y 0 , X 0 ), T 7→ T 0 , is linear and isometric, that is,
kT 0 k = kT k. In general, it is not surjective.
(ii) (ST )0 = T 0 S 0 for S ∈ L(Y, Z) and T ∈ L(X, Y ).
Chapter 2. Bounded maps; the dual space 39
kT 0 y 0 k = ky 0 ◦ T k ≤ ky 0 k kT k, y0 ∈ Y 0 ,
L : `p → `p , L(x1 , x2 , x3 , . . . ) = (x2 , x3 , . . . )
JX (x) : X 0 → K, JX (x)x0 := x0 x
JX : X → X 00 , JX (x)x0 = x0 x, x0 ∈ X 0
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 0 such that ϕx (x) = kxk (Corol-
lary 2.17). It follows that kJX (x)ϕx k = |ϕx (x)| = kxk, hence kJX (x)k ≥ 1.
The preceding theorem gives another easy proof that every normed space X can
be completed (see Theorem 1.7).
40 2.4. The Banach space adjoint and the bidual
T
X Y
JX JY
X 00 Y 00
T 00
If X and Y are identified with subspaces of X 00 and Y 00 via the canonical maps
JX and JY , then T 00 is an extension of T . Note that with this identification
S ∈ L(Y 0 , X 0 ) is adjoint operator of some T ∈ L(X, Y ) if and only if S 0 (X) ⊆ Y .
0
Lemma 2.34. Let X be a normed space. Then JX ◦ JX 0 = idX 0 .
Theorem 2.35. (i) Every closed subspace of a reflexive normed space is re-
flexive.
(ii) A Banach space X is reflexive if and only if X 0 is reflexive.
Chapter 2. Bounded maps; the dual space 41
Proof. (i) Let U be a closed subspace of a reflexive normed space X and let
u00 ∈ U 00 . We have to find a u ∈ U such that JX (u) = u00 . Let x000 : X 0 →
K, x000 (x0 ) = u00 (x0 |U ). Obviously, x000 is linear and bounded because
w
Notation: xn −
→ x or w- lim xn = x.
n→∞
42 2.4. The Banach space adjoint and the bidual
Remarks 2.38. (i) If the weak limit of a sequence exists, then it is unique,
because, by the Hahn-Banach theorem, the dual space separates points (Corol-
lary 2.17).
(ii) Every convergent sequence is weakly convergent with the same limit.
(iii) A weakly convergent sequence is not necessarily convergent. Consider for
example the sequence of the unit vectors (en )n∈N in c0 . Let ϕ ∈ c00 ∼
= `1 . Then
lim ϕ(en ) = 0 but the sequence of the unit vectors does not converge in norm.
n∈N
Example 2.39. Let (xn )n∈N be a bounded sequence in C([0, 1]). Then the
following is equivalent:
(i) (xn )n∈N converges weakly to y ∈ C[(0, 1)].
(ii) (xn )n∈N converges pointwise to y ∈ C[(0, 1)].
Proof. “(i) =⇒ (ii)” It is easy to see that for every t0 ∈ [0, 1] the point
evaluation x 7→ x(t0 ) is a bounded linear functional. Hence for all t ∈ [0, 1] the
sequence (xn (t)n∈N converges to some y(t). By assumption, [0, 1] → K, t 7→ y(t)
belongs to C([0, 1]).
“(ii) =⇒ (i)” follows from Riesz’s representation theorem (Theorem 2.24) and
the Lebesgue convergence theorem (see A.19).
Now the sequence (ϕ2 (xn1 ,j ))j∈N is bounded, so it contains a convergent subse-
quence
|x0 (yn ) − x0 (ym )| ≤ |x0 (yn ) − ϕk (yn )| + |ϕk (yn ) − ϕk (ym )| + |ϕk (ym ) − x0 (ym )|
≤ 2M kx0 − ϕk k + |ϕk (yn ) − ϕk (ym )|
ε ε
< + = ε.
2 2
This implies that (x0 (yn ))n∈N is a Cauchy sequence in K, hence it converges. To
show that (yn )n∈N converges weakly, define the map
|ψ(x0 )| = lim x0 (yn ) = lim |x0 (yn )| ≤ lim kx0 k k(yn )k ≤ M kx0 k.
n→∞ n→∞ n→∞
Now assume that X is not separable. Let Y := span{xn : n ∈ N} where (xn )n∈N
is the bounded sequence in X chosen at the beginning of the proof. Y is sep-
arable (Theorem 1.25) and reflexive (Theorem 2.35). Hence, by the first step
of the proof, there exists a subsequence (yn )n∈N ⊆ Y of (xn )n∈N and a y0 such
w
that yn −→ y0 in Y . Let x0 ∈ X 0 . Then x0 |Y 0 ∈ Y 0 , hence lim x0 (yn ) =
n→∞
w
lim x0 |Y 0 (yn ) = x0 |Y 0 (y0 ) = x0 (y0 ). Therefore we also have yn −
→ y0 in X.
n→∞
Chapter 3. Linear operators in Banach spaces 45
Chapter 3
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
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
In this way we obtain sequences (εn )n∈N and (xn )n∈N with 0 < εn < 2−n ε and
Observe that xn ∈ B(xN , ε2N ) for N ∈ N and n ≥ N . This implies that (xn )n∈N
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
n→∞
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
|fλ (x)| ≤ M, x ∈ X, λ ∈ Λ.
The next theorem shows that a family of pointwise bounded continuous func-
tions on a complete metric space is necessarily uniformly continuous on a certain
ball.
∀x ∈ X ∃ Cx ≥ 0 ∀f ∈ F kf (x)k < Cx .
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.
Let x ∈ X. Since F is pointwise bounded, there exists an nx ∈ N such that
x ∈ Anx , hence X ⊆ ∪n∈N An . By Baire’s theorem exists an N ∈ N, x0 ∈ X,
r > 0 such that Br (x0 ) ⊆ AN , that is, (3.2) is satisfied with M = N .
The Banach-Steinhaus theorem is obtained in the special case of linear bounded
functions.
∀x ∈ X ∃ Cx ≥ 0 ∀f ∈ F kf (x)k < Cx .
kf k < M, f ∈ F.
Proof. By the uniform boundedness principle there exists an open ball Br (x0 ) ⊆
X and an M 0 ∈ R such that kf (x)k < M 0 for all x ∈ Br (x0 ) and f ∈ F. For
x ∈ X with kxk = 1 and f ∈ F we find
1 1
kf (x)k = kf (rx)k = kf (x0 ) − f (x0 − rx)k
r r
1 2M 0
≤ (kf (x0 )k + kf (x0 − rx)k) ≤ =: M,
r | {z } r
∈Br (x0 )
Corollary 3.8. Let X be a normed space and A ⊆ X. Then the following are
equivalent:
(i) A is bounded.
(ii) For every x0 ∈ X 0 the set {x0 (a) : a ∈ A} is bounded.
Hence A is bounded.
48 3.2. Uniform boundedness principle
Proof. Let X be a normed space and (xn )n∈N be a weakly convergent sequence
in X. By hypothesis, for every x0 ∈ X 0 the set {x0 (xn ) : n ∈ N} is bounded.
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.
Theorem 3.10. Let (X, k · k) be a normed space, (xn )n∈N and x0 ∈ X. Then
the following is equivalent:
(i) x0 = w- lim xn .
n→∞
(ii) (xn )n∈N is bounded and there exists a total subset M 0 ⊆ X 0 such that
(i) A0 is bounded.
(ii) For all x ∈ X the set {a0 (x) : a0 ∈ A0 } is bounded.
Proof. The implication “(i) =⇒ (ii)” is clear. The other direction follows di-
rectly from the Banach-Steinhaus theorem.
Note that for “(ii) =⇒ (i)” the assumption that X is a Banach space is necessary.
For example, let d = {x = (xn )n∈N : xn 6= 0 for at most finitely many n} ⊆ `∞ .
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
delta. Obviously ϕm ∈ d0 and kϕm k = m, hence the family (ϕm ) is not bounded
in d0 , but for every fixed x ∈ d the set {ϕm (x) : m ∈ M } is.
lim kTn − T k = 0.
n→∞
s
(ii) (Tn )n∈N converges strongly to T , denoted by s- lim Tn = T or Tn −
→ T , if
n→∞
and only if
lim kTn x − T xk = 0, x ∈ X.
n→∞
Chapter 3. Linear operators in Banach spaces 49
w
(iii) (Tn )n∈N converges weakly to T , denoted by w- lim Tn = T or Tn −
→ T , if
n→∞
and only if
(ii) Convergence in norm implies strong convergence and the limits are equal.
Strong convergence implies weak convergence and the limits are equal.
The reverse implications are not true:
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 ∈
n∈N
L(X, Y ).
Theorem 3.14 (Korovkin). Let X = C[0, 2π] the space of the continuous
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
operators such that Tn xj → xj for n → ∞ and j = 0, 1, 2. Then (Tn )n∈N
converges strongly to id, that is, Tn x → x for all x ∈ X.
t−s
yt (s) = sin2 , t, s ∈ [0, 2π].
2
Now fix x ∈ X and ε > 0. Since x is uniformly continuous there exists a δ > 0
such that for all s, t ∈ [0, 2π]
t−s
yt (s) = sin2 <δ =⇒ |x(t) − x(s)| < ε.
2
2kxk∞
Setting α = δ we obtain that
|x(t) − x(s)| ≤ ε + αyt (s), s, t ∈ [0, 2π],
because either s, t are such that yt (s) < δ, then |x(t) − x(s)| < δ by definition
of δ; or yt (s) ≥ δ, then |x(t) − x(s)| ≤ 2kxk∞ = αδ ≤ αyt (s). Hence we have
that
−ε − αyt (s) ≤ x(t) − x(s) ≤ ε + αyt (s), s, t ∈ [0, 2π]
=⇒ −εx0 − αyt ≤ x(t)x0 − x ≤ εx0 + αyt , t ∈ [0, 2π]
and since Tn is positive and yt is a positive function
−εTn x0 − αTn yt ≤ x(t)Tn x0 − Tn x ≤ εTn x0 + αTn yt , t ∈ [0, 2π].
Fourier Series
Definition 3.15. Let x : R → R a 2π-periodic integrable function. The Fourier
series of x is
∞
a0 X
S(x, t) = + (ak cos(kt) + bk sin(kt)),
2
k=1
where
1 π
Z
ak := x(s) cos(ks) ds, k ∈ N0 ,
π −π
1 π
Z
bk := x(s) sin(ks) ds, k ∈ N.
π −π
Chapter 3. Linear operators in Banach spaces 51
Note that the Fourier series is a formal series only. In the following we will
prove theorems on convergence of the Fourier series.
First we will use methods from Analysis 1 to show that for a continuously differ-
entiable periodic function its Fourier series converges uniformly to the function.
Next we will use the uniform boundedness principle to show that there exist
continuous functions whose Fourier series does not converge pointwise every-
where. Finally, the Korovkin theorem implies that the arithmetic means of the
partial sums of the Fourier series of a periodic function converges uniformly to
the function.
For a given 2π-periodic function and n ∈ N we define the nth partial sum
n
a0 X
sn (x, t) = + (ak cos(kt) + bk sin(kt)). (3.3)
2
k=1
Lemma 3.16.
( sin((n+ 1 )s)
π
s 6= 0,
Z
1 2
2 sin( 2s ) ,
sn (x, t) = x(s + t)Dn (s) ds with Dn (s) =
π −π n + 12 s = 0.
(3.4)
Dn is called Dirichlet kernel. Dn is continuous and
1 π
Z
Dn (s) ds = 1. (3.5)
π −π
Proof. Using the trigonometric identity cos(a) cos(b) + sin(a) sin(b) = cos(a − b)
and that x is 2π-periodic we obtain
n
a0 X
sn (x, t) = + (ak cos(kt) + bk sin(kt))
2
k=1
n
1 π
Z 1 X
= x(s) + (cos(ks) cos(kt) + sin(ks) sin(kt)) ds
π −π 2
k=1
Z π n
1 1 X
= x(s) + cos(k(s − t)) ds
π −π 2
k=1
n
1 π
Z 1 X
= x(s + t) + cos(ks) ds.
π −π 2
k=1
1 1
Pn
Note that lim Dn (s) = n + 2 = 2 + k=1 cos(0). For the proof of (3.5) let
s→0
x = 1 a constant function on R. Then, by (3.3),
1 π
Z
Dn (s) ds = sn (x, t) = x(t) = 1.
π −π
We have to show that An (t), Bn (t) and Cn (t) tend to 0 for n → ∞ uniformly
in t. Using the mean value theorem and that π2 σ ≤ sin(σ) for σ ∈ [0, π/2] we
obtain
Z h Z h
|x(s + t) − x(t)| 1 kx0 k |s|
Bn (t) = s | sin((n + 2 )s)| ds ≤ s ds
−h 2 sin | 2 | | {z } −h 2 sin | 2 |
≤1
π0ε
≤ 2hkx k∞ < .
2 2
Define the auxiliary function
x(s + t) − x(t)
ft (s) = , s ∈ [h, π], t ∈ [0, π].
2 sin( 2s )
2kxk∞
The functions ft are continuously differentiable and kft k∞ ≤ 2 sin(h/2) =: M1 ,
0
kx k∞
kft0 k∞
≤ =: M2 . Note that the bounds do not depend on t. Integrating
2 sin(h/2)
by parts, we find
Z π
Cn (t) = ft (s) sin((n + 21 )s) ds
h
cos((n + 1 )s) π Z π cos((n + 1 )s)
2 2 0
=− f (s) + f (s) ds
1 t 1 t
n+ 2 h h n+ 2
1 M
≤ 1 (2M1 + (π − h)M2 ) =: .
n+ 2 n + 21
Finally we show that the arithmetic mean of the partial sums of the Fourier
series of a continuous function converge.
Proof. Note that the Tn are well-defined and that for all x ∈ X and t ∈ [−π, π]
n−1 Z n−1
1X π
Z π
1 x(s + t)) X 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
we can write Tn x as
Z
1
Tn x(t) = −π π Fn (s)x(s + t) ds.
π
Note that all Fn are positive functions, hence the Tn are positive operators. To
show the theorem, it suffices to show that Tn xj → xj for x0 (t) = 1, x1 (t) =
cos(t), x2 (t) = sin(t) (Korovkin theorem). Using (3.3) it follows that sk (x0 , · ) =
x0 for all k ∈ N0 and that
s0 (x1 , · ) = s0 (x2 , · ) = 0,
sk (x2 , · ) = x1 , sk (x1 , · ) = x2 , k ∈ N.
n−1
Since Tn x0 = x0 , Tn xj = n xj for j = 1, 2 and n ∈ N the theorem is proved.
Chapter 3. Linear operators in Banach spaces 55
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
A := {(s, t) ∈ R × R : s ≥ 0, st ≥ 2} is closed in R × R but π(A) = (0, ∞) is
open in R.
The lemma says that if T (BX (0, 1)) is dense in BY (0, r), then, for any 0 < ρ < r,
the ball BY (0, ρ) is contained in T (BX (0, 1)).
Proof. Note that the assertion is equivalent to
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)).
Hence there exists an x1 ∈ BX (0, 1) such that ky0 − T x1 k < εr. By scaling, we
know that T (BX (0, ε)) is dense in BY (0, εr). Since y0 − T x1 ∈ BY (0, εr), there
exists an x2 ∈ BX (0, ε) such that ky0 − T x1 − T x2 k < ε2 r. Since T (BX (0, ε2 ))
is dense in BY (0, ε2 r), there exists an x3 ∈ BX (0, ε2 ) such that ky0 − T x1 −
T x2 − T x3 k < ε3 r.
Continuing in this way, we obtain a sequence (xn )n∈N such that
n
X
kxn k < εn−1 , ky0 − T xk k < rεn , n ∈ N. (3.6)
k=1
P∞ P∞
It
Pfollows that x0 := k=1 xk exists and lies in B(0, (1−ε)−1 ) because k=1 kxk k <
∞
k=1 rε
k−1
= r(1 − ε)−1 . Since T is continuous, we know that
∞
X ∞
X
T (x0 ) = T xk = T xk .
k=1 k=1
Pn
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
In the proof of the open mapping theorem we use the following fact.
Proof. Obviously it suffices to show that T (BX (0, 2)) is dense in BY (0, 2δ).
Since T is linear, it follows immediately that TX (B(0, 1)) is dense in BY (−y, δ).
Let z ∈ BY (0, 2δ) and ε > 0. Note that y − z/2 ∈ BY (y, δ) and −y − z/2 ∈
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
it follows that z ∈ T (BX (0, 2)) because ε can be chosen arbitrarily small.
G(T ) := {(x, T x) : x ∈ D} ⊆ X × Y
Proof. Assume that T is closed and let (xn )n∈N such that (xn )n∈N and (T xn )n∈N
converge. Then ((xn , T xn ))n∈N ⊆ G(T ) converges in X × Y . Since G(T ) is
closed, lim (xn , T xn ) = (x0 , y0 ) ∈ G(T ). By definition of G(T ) this implies
n→∞
lim xn = x0 ∈ D(T ) and T x0 = y0 = lim T xn .
n→∞ n→∞
Now assume that (3.7) holds and let ((xn , T xn ))n∈N ⊆ G(T ) be a sequence
that converges in X × Y . Then both (xn )n∈N and (T xn )n∈N converge, hence
x0 := lim xn ∈ D and lim T xn = T x0 which shows that lim (xn , T xn ) =
n→∞ n→∞ n→∞
(x0 , T x0 ) ∈ G(T ), hence G(T ) is closed.
Proof. Assume that T is closable. Then G(T ) is the graph of a linear function.
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
is linear.
Now assume that (3.8) holds and define T as in (3.9). T is well-defined because
for sequences (xn )n∈N and (x̃n )n∈N in D with lim xn = lim x̃n = x such that
n→∞ n→∞
(T xn )n∈N and (T x̃n )n∈N in D converge, it follows that (xn − x̃n )n∈N converges
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
n→∞ n→∞ n→∞
definition, G(T ) is the closure of G(T ), so T is the closure of T .
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-
n→∞
entiable and T x0 = x00 = ( lim xn )0 = lim x0n = lim T xn .
n→∞ n→∞ n→∞
That T is not continuous was already shown in Example 2.7 (iv) (choose
xn (t) = n1 exp(−n(t + 1))).
Proof. By Lemma 3.32 and the open mapping theorem (Theorem 3.22) the
operator iTe : (D, k · kT ) → Y, Tex = T x, is open. Let U ⊆ D open with respect
to the norm in X. Then U is also open with respect to the graph norm because
obviously i : (D, k · kT ) → (D, k · k), ix = x, is bounded, hence continuous.
Hence T (U ) = Te(U ) is open in Y .
Now assume in addition that T is injective. Then Te−1 : Y → (D, k · kT ) is
continuous by the inverse mapping theorem. Since i is continuous, also T −1 =
(Te ◦ i−1 )−1 = i ◦ Te−1 is continuous.
Proof. Let (yn )n∈N be a Cauchy sequence in rg(T ) with y0 := lim yn . and
n→∞
xn := T −1 yn , n ∈ N. Then (xn )n∈N is a Cauchy sequence in D because kxn −
xm k = kT −1 yn − T −1 ym k ≤ kT −1 k 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
n→∞
closed.
Proof. (i) =⇒ (ii) follows from the closed graph theorem because by assumption
D is Banach space.
(ii) =⇒ (iii) and (iii) =⇒ (i) are clear.
Let X be an infinite dimensional Banach space and (xλ )λ∈Λ an algebraic basis
of X. Without restriction we can assume kxλ k = 1, λ ∈ Λ. Choose N → Λ, n 7→
λn be an injection. Then the operator
X X
T : X → X, T (x) = n cλn xλn for x = cλn xλn ∈ X,
n∈N λ∈Λ
Lemma 3.39. Let X be a normed space and P ∈ L(X) a projection. Then the
following holds:
(i) Either P = 0 or kP k ≥ 1.
(ii) ker(P ) and rg(P ) are closed.
(iii) X is isomorphic to ker P ⊕ rg(P ).
Proof. From linear algebra we know that there exist bases (u1 , . . . , un ) of U and
(ϕ1 , . . . , ϕn ) of U 0 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
on X with kϕk k = kψk k. We define
n
X
P : X → X, Px = ϕk (x)uk .
k=1
Proof. (i) Since U and V are Banach spaces, their sum U ⊕V is a Banach space.
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.
(ii) P : X → U, u + v 7→ u is the desired projection.
(iii) The map V 7→ X/V, v 7→ [v] is linear, bijective and continuous. Since U is
closed, X/U is a Banach space. By the inverse mapping theorem it follows that
V and X/U are isomorphic.
Remark 3.43. Note that not every closed subspace of a Banach space is com-
plemented in the sense of the theorem above. For example, c0 is not comple-
mented as subspace of `∞ .
Chapter 3. Linear operators in Banach spaces 63
Proof. Let τ (U) be the topology generated by U and σ(U) the system of sets
described in (3.10). It is not hard to see that σ(U) is a topology containing U,
hence containing τ (U). On the other hand, all sets of the form (3.10) are open
in τ (U), so σ(U) ⊆ τ (U).
Definition 3.46. Let X be a set, Λ be an index set and for every λ ∈ Λ let
(Yλ , τλ ) be a topological space. Consider a family F = (fλ : X → Yλ ) of
functions. The smallest topology on X such that all fλ are continuous, is called
the initial topology on X, denoted by σ(X, F).
Note that τ (F) = τ {fλ−1 (Uλ ) : λ ∈ Λ, Uλ ∈ τλ }.
Proof. Assume that (xn )n∈N is weakly convergent with x0 := w- lim xn and let
n→∞
U be a σ(X, X 0 )-open set containing x0 . Then there exist ϕ1 , . . . , ϕn such that
n
\
x0 ∈ ϕ−1
j (Vj ) ⊆ U
k=1
with Vj open subsets in R containing ϕj (x0 ). Since lim ϕ(xn ) = ϕ(x0 ) for all
n→∞
64 3.6. Weak convergence
Lemma 3.49. Let X be a normed space, (xn )n∈N ⊆ X and (ϕn )n∈N ⊆ X 0 .
(i) x0 = w- lim xn =⇒ kx0 k ≤ lim inf xn .
n→∞ n→∞
Proof. (i) For x0 = 0 the assertion is clear. By the Hahn-Banach theorem there
exists an ϕ ∈ X 0 such that ϕ(x0 ) = kx0 k and kϕk = 1. Hence
kx0 k = k lim ϕ(xn )k ≤ lim inf kϕk kxn k = lim inf kxn k.
n→∞ n→∞ n→∞
(ii) Let ε > 0. Then there exists an x ∈ X with kxk = 1 such that kϕ0 k − ε <
kϕ0 (x)k. The statement follows as above:
kϕ0 k − ε < kϕ0 (x)k = lim kϕn (x)k ≤ lim inf kϕn k kxk = lim inf kϕn k.
n→∞ n→∞ n→∞
Hence the lemma above states that k · k is lower semicontinuous in the weak
topology.
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
Chapter 3. Linear operators in Banach spaces 65
U := {a ∈ A : |a(x + y) − ϕ(x + y)| < ε, |a(x) − ϕ(x)| < ε, |a(y) − ϕ(y)| < ε}
Since ε was arbitrary, this implies ϕ(x + y) = ϕ(x) + ϕ(y). Similarly it can be
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 ∈ K10 .
Chapter 4. Hilbert spaces 67
Chapter 4
Hilbert spaces
h· , ·i : X × X → K
• hermitian ⇐⇒ hx , yi = hy , xi, x, z ∈ X,
• positive semidefinite ⇐⇒ hx , xi ≥ 0, x ∈ X,
• positive (definite) ⇐⇒ hx , xi > 0, x ∈ X \ {0}.
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 +
λyi = 0 with λ chosen as above, so x and y are linearly dependent.
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.
Lemma 4.4. An inner product space (X, h· , ·i) becomes a normed space by
1
setting kxk := hx , xi 2 , x ∈ X.
Proof. The only property of a norm that does not follow immediately from the
definition of k · k is the triangle inequality. To prove the triangle inequality,
choose x, y ∈ X. Using the Cauchy-Schwarz inequality, we find
In the following, we will always consider inner product spaces endowed with the
topology induced by the norm.
Lemma 4.6. Note that the scalar product on a inner product space X is a
continuous map X × X → K when X × X is equipped with the norm k(x, y)k =
kxkX + kykX .
The polarisation formula allows to express the inner product of two elements of
X in terms of their norms.
Chapter 4. Hilbert spaces 69
1
hx , yi = kx + yk2 − kx − yk2 , if K = R,
4
1
hx , yi = kx + yk2 − kx − yk2 + ikx + iyk2 − ikx − iyk2 , if K = C.
4
Proof. Assume that the norm is generated by the inner product h· , ·i and let
1
kxk = hx , xi 2 . Then for all x, y ∈ X parallelogram identity holds:
Now assume that the norm on X is such that the parallelogram identity holds
and for x, y ∈ X define hx , yi by the polarisation formula. We prove that h· , ·i
is an inner product on X in the case K = C. The case K = R can be proved
analogously.
• Positivity.
• Hermiticity.
• Additivity.
4(hx , yi + hx , zi)
= kx + yk2 − kx − yk2 + i kx + iyk2 − i kx − iyk2
+ kx + zk2 − kx − zk2 + i kx + izk2 − i kx − izk2
y+z y − z
2
y+z y − z
2
=
x + +
−
x − −
2 2 2 2
y+z y − z
2
y+z y − z
2
+
x + −
−
x − +
2 2 2 2
y+z y − z
2
y+z y − z
2
+ i
x + i +i
− i
x − i −i
2 2 2 2
y+z y − z
2
y+z y − z
2
+ i
x + i −i
− i
x − i +i
2 2 2 2
y + z
2
y − z
2 y + z
2
y − z
2
= 2
x +
+ 2
− 2
x −
− 2
2 2 2 2
y − z
2
y + z
2
y + z
2
y − z
2
+ 2i
x + i
+ 2i
− 2i
x − i
− 2i
2 2 2 2
y + z
2 y + z
2 y + z
2
2
y + z
= 2
x +
− 2
x −
+ 2i
x + i
− 2i
x − i
2 2 2 2
y+z
= 2 · 4hx , i.
2
If we choose z = 0 we find hx , yi = 2hx , y2 i, hence
y + z
hx , yi + hx , zi = 2 x , = hx , y + zi.
2
Proof. By continuity of the norm, the parallelogram identity holds on the com-
pletion X of an inner product space X. So X is an inner product space.
(iii) Let R([0, 1]) be the vector space of the Riemann integrable functions on
the interval [0, 1]. Then
Z 1
hf , gi = f (t)g(t) dt, f, g ∈ R([0, 1]),
0
defines a sesquilinear form on R([0, 1]) which is not positive definite, since,
for example, χ{0} 6= 0, but hχ{0} , χ{0} i = 0.
The restriction of h· , ·i to the space of the continuous functions C([0, 1]) is
an inner product which is not complete (its closure is the space L2 ([0, 1])).
4.2 Orthogonality
Definition 4.11. Let X be an inner product space.
M ⊥ := {x ∈ X : x ⊥ m, m ∈ M }.
Lemma 4.14. Let M be a closed and convex subset of a Hilbert space H and
fix x0 ∈ H. For y0 ∈ M the following are equivalent:
(i) kx0 − y0 k = dist(x0 , M ),
(ii) Rehx0 − y0 , y − y0 i ≤ 0, y ∈ M.
2 Rehx0 − y0 , y − y0 i ≤ tky − y0 k2
Proof of Theorem 4.16. Fix x0 ∈ H and let PU (x0 ) := y0 the unique element
y0 ∈ U such that kx0 − y0 k = dist(x0 , U ). Then rg(PU ) = U and PU2 = PU ,
hence PU is a projection on U .
By Lemma 4.15, PU (x0 ) is the unique element in U such that x0 −PU (x0 ) ∈ U ⊥ .
Rehx0 − PU (x0 ) , y − PU (x0 )i ≤ 0, y ∈ U.
Proof. By the projection theorem (Theorem 4.16), for every closed subspace V
Φ : H → H 0, y 7→ h· , yi
hence kΦ(y)k ≤ kyk If y 6= 0, then set x = kyk−1 y. Note that kxk = 1 and
kΦ(y)xk = kyk, implying that kΦ(y)k = kyk. So we have shown that Φ is
well-defined and an isometry. In particular, Φ is injective.
To show that Φ is surjective, fix an ϕ ∈ H 0 . If ϕ = 0, then ϕ = Φ(0). Otherwise
we can assume that kϕk = 1. Since ker{ϕ} is closed, there exists a decomposition
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 ∈
ker ϕ ⊕ (ker ϕ)⊥ ,
hxn − x0 , yi → 0, y ∈ H.
(ii) Every bounded sequence (xn )n∈N ⊆ H contains a weakly convergent sub-
sequence.
Proof. (i) follows from the Riesz-Fréchet theorem, and (ii) follows with Theo-
rem 2.40.
S ⊆ T =⇒ S = T.
Examples 4.24. (i) The unit vectors (en )n∈N in `2 (N) are a orthonormal
system.
Proof. Let s1 := kx1 k−1 x1 . Next set y2 := x2 − hx1 , s1 is1 . Note that y2 6= 0
because x2 and x1 are linearly independent. Let s2 := ky2 k−1 y2 . Then s1 ⊥ s2
and ks1 k = ks2 k = 1. Now for k ≥ 1 let
n
X
yn+1 := xn+1 − hxk , sk isk , sn+1 := kyn+1 k−1 yn+1 .
k=1
PN
Proof. For N ∈ N let xN := x− n=1 hx , sn isn . Since xN ⊥ sn for n = 1, . . . , N ,
Pythagoras’ theorem yields
XN
2 N
X N
X
kxk2 = kxN k2 +
hx , sn isn
= kxN k2 + |hx , sn i|2 ≥ |hx , sn i|2 .
n=1 n=1 n=1
Sx := {λ ∈ Λ : hx , sλ i =
6 0}
is at most countable.
Now we show that the orthonormal systems in Example 4.24 are complete.
Examples 4.32. (i) The set of the unit vectors {en : n ∈ N} in `2 (N) are a
complete orthonormal system in `2 (N) because {en : n ∈ N} = `2 (N).
P
Then hf , gi = γ∈Γ f (γ)g(γ) is a well-defined inner product (note that
only countably many terms are 6= 0 and the sum is absolutely convergent
by Hölder’s inequality). As in the case Γ = N it can be shown that `2 (Γ)
is a Hilbert space and (fλ )λ∈Γ where fλ (γ) = δλγ (Kronecker delta) is a
complete orthonormal system in `2 (Γ).
Chapter 4. Hilbert spaces 79
Lemma 4.33. Let H be an infinite dimensional Hilbert space. Then the fol-
lowing is equivalent.
(i) H is separable.
(ii) Every complete orthonormal system in H is countable.
(iii) There exists an countable complete orthonormal system in H.
Proof. The statement is proved in linear algebra if |S| < ∞. Now assume that
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|.
Analogously, |S| ≤ |T ||N| = |T |. By the Schröder-Bernstein theorem then
|S| = |T |.
Hence T ∗ is characterised by
hT x , yi = hx , T ∗ yi, x ∈ H1 , y ∈ H2 .
(i) (λS + T )∗ = λS ∗ + T ∗ .
(ii) (RT )∗ = T ∗ R∗ .
(iii) T ∗ ∈ L(H2 , H1 ) and kT ∗ k = kT k.
(iv) T ∗∗ = T .
(v) kT T ∗ k = kT ∗ T k = kT k2 .
(vi) ker T = (rg(T ∗ ))⊥ , ker T ∗ = (rg(T ))⊥ .
(vii) If T is invertible, then (T −1 )∗ = (T ∗ )−1 .
Proof. (i)–(iv) are clear. For the proof of (v) note that for kxk = 1
kT xk2 = hT x , T xi = hx , T ∗ T xi ≤ kxk kT ∗ T xk ≤ kT ∗ T k ≤ kT ∗ k kT k = kT k2 .
Taking the supremum over all x ∈ H with kxk = 1 shows the desired equalities.
Tx = 0 ⇐⇒ ∀y ∈ H2 hT x , yi = 0 ⇐⇒ ∀y ∈ H2 hx , T ∗ yi = 0
∗
⇐⇒ x ⊥ rg(T ).
Next we show that a length preserving linear map between Hilbert spaces also
preserves angles.
Proof. (i) The direction “⇐” is clear; “⇒” follows from the polarisation formula
(Theorem 4.7).
(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
hT x , T yi = hT ∗ T x , yi = hx , yi,
hx , y − T ∗ T yi = hx , yi − hT x , T yi = 0, x, y ∈ H1 ,
hT ∗ ξ , T ∗ ηi = hT ∗ T x , T ∗ T yi = hx , yi = hT x , T yi = hξ , ηi.
that is Tk∗ = Tk .
82 4.4. Linear operators in Hilbert spaces
hT x , yi = hx , T yi, x, y ∈ H.
Theorem 4.44. Let H be a complex Hilbert space. For T ∈ L(H) the following
is equivalent.
(i) hT x , xi ∈ R, x ∈ H.
(ii) T is selfadjoint.
hT x , xi = hx , T xi = hT x , xi, x ∈ H.
hT x , yi + hT y , xi = hy , T xi + hx , T yi,
hT x , yi − hT y , xi = −hy , T xi + hx , T yi,
so finally hT x , yi = hx , T yi.
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.
Chapter 4. Hilbert spaces 83
1
RehT x , yi ≤ |hT (x + y) , x + yi| + |hT (x − y) , x − yi|
4
1 M
≤ M kx + yk2 + M kx − yk2 = (kxk2 + kyk2 ) ≤ M.
4 2
Now choose λ ∈ C, |λ| = 1 such that λhT x , yi = |hT x , yi|, so
kT xk = kT ∗ xk, x ∈ H,
hP x , yi = hP x , y − P y + P yi = hP x , P yi,
hx , P yi = hx − P x + P x , P yi = hP x , P yi.
it follows that hx , yi = 0.
P1 P2 + P2 P1 = (P1 + P2 )2 − (P1 + P2 ) = 0.
In particular 0 = (P1 P2 +P2 P1 )P2 x = (id +P2 )P1 P2 x. Note that for y ∈ H \{0}
the vectors (id −P2 )y and P2 y are linearly independent, hence (id +P2 )y =
(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}.
(i) H0 ⊆ H1 ,
(ii) kP0 xk ≤ kP1 xk, x ∈ H.
(iii) hP0 x , xi ≤ hP1 x , xi, x ∈ H.
(iv) P0 P1 = P0 .
(iv) =⇒ (ii) For all x ∈ H: kP0 xk = kP0 P1 xk ≤ kP0 kkP1 xk ≤ kP1 xk.
(iii) =⇒ (i) Let x ∈ H1⊥ = ker P1 . Then 0 = hP1 x , xi ≥ hP0 x , xi = kP0 k2 ,
hence H1⊥ ⊆ ker P0 = H0⊥ .
Lemma 4.55. Let H be a Hilbert space and (Pn )n∈N a sequence of orthogonal
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.
Chapter 4. Hilbert spaces 87
P 2 x = (P − Pn + Pn )(P − Pn + Pn )x = (P − Pn )P x + Pn (P − Pn )x + Pn2 x.
is well-defined.
(i) Let (S, D(S)) and (T, D(T )) be densely defined linear operators X → Y .
If S ⊆ T then T 0 ⊆ S 0 .
(ii) Assume S(X → Y ) and T (Y → Z) are densely defined such that also T S
is densely defined. Then S 0 T 0 ⊆ (T S)0 .
(iii) Assume S(X → Y ) and T (X → Y ) are densely defined such that also
T + S is densely defined. Then (S 0 + T 0 ) ⊆ (S + T )0 .
Remark 4.62. The sets A◦ and ◦ B are closed subspaces and ◦ (A◦ ) = A. If X
is reflexive, then also (◦ B)◦ = B.
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 ∈
(X × Y )0 such that ψ|G(T ) = 0 and ψ((0, y0 )) 6= 0. Let ϕ : Y → K, ϕ(y) =
ψ((0, y)). Obviously ϕ ∈ Y 0 and ϕ(y0 ) 6= 0. Moreover, ϕ ∈ D(T 0 ) because for
all x ∈ D(T )
ϕ(T x) = ψ((0, T x)) = ψ((x, T x) − (x, 0)) = ψ((x, T x)) − ψ((x, 0))
= −ψ((x, 0)).
Theorem 4.64. Let X and Y be Banach spaces. For a densely defined closed
linear operator T (X → Y ) the following holds:
◦
(i) rg(T )◦ = rg(T ) = ker T 0 .
(ii) rg T = ◦ (ker T 0 ).
(iii) rg T = Y ⇐⇒ T 0 is injective.
(iv) ◦ (rg T 0 ) ∩ D(T ) = ker T.
(v) rg T 0 ⊆ (ker T )◦ .
Proof. (i) The first equality is clear. The second equality follows from
(iii) By (ii), rg T = Y if and only if ◦ (ker T 0 ) = Y . This is the case if and only
if ϕ(y) = 0 for all ϕ ∈ ker T 0 and y ∈ Y , that is, if and only if ker T 0 = {0}.
(iv) Let x ∈ ker(T ) and x0 ∈ rg T 0 . Choose y 0 ∈ D(T 0 ) with T 0 y 0 = x0 . Then
x0 x = (T 0 y 0 )x = y 0 (T x) = y 0 (0) = 0, hence x ∈ ◦ (rg T 0 ).
Now let x ∈ ◦ (rg T 0 ) ∩ D(T ). Then y 0 (T x) = (T 0 y 0 )x = 0 for all y 0 ∈ Y 0 . Since
T is closed, it follows by Lemma 4.63 that T x = 0, hence x ∈ ker T .
(v) Let x0 ∈ rg(T 0 ) and x ∈ ker T . Choose y 0 ∈ D(T 0 ) such that T 0 y 0 = x0 .
Then x0 x = (T 0 y 0 )x = y 0 (T x) = y 0 (0) = 0. It follows that rg(T 0 ) ⊆ (ker T )◦ ,
and since (ker T )◦ is closed, the statement is proved.
(T 0 )−1 = (T −1 )0 (4.2)
Proof.
Te : (D(T ), k · kT ) → rg T, Tex = T x
(iii) ⇐⇒ (iv) Recall that T is open if and only if there exists an , r > 0 such
that the image of the open ball in X with centre 0 and radius r contains the open
unit ball in Y . That is, there exists a r > 0 such that T (BX (0, r)) ⊇ BY (0, 1).
Assume that T is open and let r as above.
To show that T 0 is open, we have to show that for every x00 ∈ rg(T 0 ) with
kx00 k < 1, there exists a y00 ∈ D(T 0 ) with T 0 y00 = x00 and ky00 k < r. Define
a linear functional ϕ on rg(T ) as follows: for y ∈ rg T with kyk < 1 choose
x ∈ D(T ) such that kxk < r and T x = y. Set ϕ(y) = x00 x and extend ϕ linearly
to rg T . Note that |ϕ(y)| = |x00 x| ≤ kx00 kkxk ≤ rkyk, ϕ is bounded, so by
the theorem of Hahn-Banach it can be extended to a functional y00 ∈ Y 0 with
ky00 k ≤ r. Note that
Note that the formal adjoint of a non-densely defined linear operator is not
unique; in particular, the operator trivial operator with D = {0} is formally
adjoint to every linear operator.
If T is densely defined, then its adjoint T ∗ is its maximal formally adjoint
operator.
92 4.6. The adjoint of an unbounded operator
Proof. Observe that U is unitary, hence U (G(T )⊥ ) = [U (G(T ))]⊥ . The first
equality in (4.3) follows from
Theorem 4.70. Let H1 and H2 be Hilbert spaces. For a densely defined linear
operator T (X → Y ) the following holds:
(i) T ∗ is closed.
(ii) If T is closable, then T ∗ is densely defined and T ∗∗ = T .
0 = h(0, y0 ) , (−z, y)iH1 ×H2 = h(0, y0 ) , U (y, z)iH1 ×H2 , (y, z) ∈ G(T ∗ ).
G(T ∗∗ ) = [V (G(T ∗ ))]⊥ = [V U (G(T )⊥ )]⊥ = [−(G(T )⊥ )]⊥ = G(T )⊥⊥ = G(T )
= G(T ).
hence T ∗∗ = T .
(T ∗ )−1 = (T −1 )∗ =: T −∗ .
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.
⊥⊥
(ii) By (i) rg(T ) = rg(T ) = (ker T ∗ )⊥ .
(iii) By Theorem 4.70 T ∗ is closed and densely defined and T ∗∗ = T . Appli-
cation of (i) to T ∗ shows rg(T ∗ )⊥ = ker T .
(iv) Application of (ii) to T ∗ shows rg(T ∗ ) = (ker T )⊥ .
For k = 1, 2, 3 let
Tk : H ⊇ D(Tk ) → H, Tk x = ix0 .
In particular we obtain
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.
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}.
σ(T ) = σp (T ) ∪˙ σc (T ) ∪˙ σr (T ).
Nλ (T ) := ker(T − λ),
Aλ (T ) := {x ∈ X : (T − λ)n x = 0 for some n ∈ N}.
Remark 5.4. Often the resolvent set of a linear operator is defined slightly
differently: Let T (X → X) is a densely defined linear operator. Then λ ∈ ρ(T )
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
following cases holds:
(i) λ − T is not bijective =⇒ λ ∈ / ρ(T );
(ii) λ − T is bijective, then (λ − T )−1 is defined everywhere and not closed, so
it cannot be bounded, which implies λ ∈ / ρ(T ).
Proof. Let λ ∈ C. Then there exists a polynomial Q such that P (X) − P (λ) =
(X − λ)Q(X). In particular, P (T ) − P (λ) = (T − λ)Q(T ) = Q(T )(T − λ).
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 )).
Now assume µ ∈ σ(P (T )). There exist a, λ1 , . . . , λn ∈ C such that P (X) − µ =
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
spectrum of T and µ = P (λj ) ∈ P (σ(T )).
Note that (ii) shows that locally around a λ0 ∈ ρ(T ) the resolvent has a power
series expansion with coefficients depending only on λ0 and T .
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-
mann series (Theorem 2.10)
∞
X
(id −S)−1 = Sn.
n=0
If |λ0 − λ| < k(λ0 − T )−1 k−1 , then the term in brackets is invertible, hence so
is λ − T and we obtain
−1
(λ − T )−1 = (λ0 − T )−1 id −(λ0 − λ)(λ0 − T )−1
X∞
−1
= (λ0 − T ) (λ0 − λ)n (λ0 − T )−n
n=0
∞
X
= (λ0 − λ)n (λ0 − T )−(n+1)
n=0
which proves (ii). If µ ∈ C with |µ| < k(T − λ0 )−1 k−1 , then λ0 + µ ∈ ρ(T ),
hence dist(λ0 , σ(T )) ≥ k(T − λ0 )−1 k−1 , so also (i) is proved.
As a corollary we obtain the following theorem.
= (µ − λ)R(λ, T )R(µ, T ).
f (z) − f (z0 )
lim
z→z0 z − z0
exists in the norm topology. f is called holomorphic if and only if it is
holomorphic in every z0 ∈ Ω.
(ii) f is called weakly holomorphic in z0 ∈ Ω if and only if the limit
f (z) − f (z0 )
lim
z→z0 z − z0
exists in the weak topology. f is called weakly holomorphic if and only if
it is weakly holomorphic in every z0 ∈ Ω. Hence, for every ϕ ∈ X 0 the
map Ω → C, z 7→ ϕ(f (z)) is holomorphic in the usual sense.
Chapter 5. Spectrum of linear operators 101
Proof. Assume that (xn )n∈N ⊆ X is a Cauchy sequence and let ε > 0. Then
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 0 with
kϕk ≤ 1.
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 0 with kϕk ≤ 1. Recall that the map
JX : X → X 00 is an isometry. It follows for m, n ≥ N
where Γr (z0 ) is the positively oriented boundary of Kr (z0 ). More generally, for
n ∈ N0 ,
Z
(n) n! f (z)
f (a) = dz, a ∈ Br (z0 ). (5.2)
2πi Γr (z0 ) (z − a)n+1
For a ∈ Br (z0 ) and 0 < |h| < r − |z0 − a| it follows that a + h ∈ Kr (z0 ), hence
102 5.2. The resolvent
1
ϕ(f (a + h)) − ϕ(f (a)) − (ϕ ◦ f )0 (a)
h Z
1 1h 1 1 h i
= − − ϕ(f (z)) dz
2πi Γr (z0 ) h z − a − h z − a (z − a)2
Z
1 h 1 1 i
= − ϕ(f (z)) dz
2πi Γr (z0 ) (z − a)(z − a − h) (z − a)2
Z
h ϕ(f (z))
= dz.
2πi Γr (z0 ) (z − a)2 (z − a − h)
Hence we obtain
1 d
ϕ(f (a + h)) − ϕ(f (a)) − (ϕ ◦ f )(a) ≤ hkϕkC 0 .
h dz
Proof. (i) =⇒ (ii) follows from the definition. (ii) ⇐⇒ (iii) follows form Theo-
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 0
Z
1 d h ϕ(T (z)x)
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)
Chapter 5. Spectrum of linear operators 103
is holomorphic.
Proof. Let λ0 ∈ ρ(T ) and λ ∈ C with |λ − λ0 | < kR(λ0 , T )k. For fixed x ∈ X
and ϕ ∈ X 0 we have by Lemma 5.7
∞
X
ϕ(R(λ, T )x) = ϕ (λ − λ0 )n (R(λ0 , T ))n+1 x
n=0
∞
X
(λ − λ0 )n ϕ (R(λ0 , T ))n+1 x
=
n=0
where we used that the operator series converges and ϕ is continuous. Since
the last sum is absolutely convergent, it follows that λ → ϕ(R(λ, T )x) is ana-
lytic locally at λ0 , hence holomorphic. Since weak holomorphy is equivalent to
holomorphy in the operator norm (Theorem 5.14), the theorem is proved.
104 5.2. The resolvent
Proof. Assume σ(T ) = ∅. Observe that this implies X 6= {0} and T −1 ∈ L(X).
Let λ ∈ C with |λ| > kT k. Then λ ∈ ρ(T ) and using the Neumann series
∞ ∞
X
X 1
kR(λ, T )k =
λn T −(n+1)
≤ |λn |kT k−(n+1) = .
n=0 n=0
kT k − |λ|
Proof. (i) Obviously, T is unbounded and densely defined. If (xn )n∈N ⊆ D(T )
such that xn → x and T xn → y ∈ X, then, by a theorem of Analysis 1, x is
differentiable, hence in D(T ) and T x = x0 = y which implies that T is closed.
For every λ ∈ C the differential equation x0 −λx = 0 has the solution xλ (t) = eλt .
Note that xλ ∈ D(T ) and (T − λ)xλ = 0, so λ ∈ σp (T ).
(ii) Obviously, T is unbounded and densely defined. If (xn )n∈N ⊆ D(T ) such
that xn → x and T xn → y ∈ X, then, by a theorem of Analysis 1, x is
differentiable and x0 = y. Moreover, x(0) = lim xn (0) = 0, so in D(T ) and
n→∞
T x = x0 = y which implies that T is closed.
For every λ ∈ C and every y ∈ X the initial value problem x0 − λx = y, x(0)
has exactly one solution xλ given by
Z t
xλ (t) = eλt e−λs y(s) ds.
0
Chapter 5. Spectrum of linear operators 105
Obviously xλ ∈ C 1 [0, 1], xλ (0) = 0 and x0λ (0) = λxλ (0) + y(0) = 0. Hence T − λ
is bijective, in particular λ ∈ ρ(T ).
Note that in the last example the continuity of (T − λ) can be seen immediately:
n Z t o
−1
e−λs y(s) ds : t ∈ [0, 1]
λt
k(T − λ) yk∞ = kxλ k∞ = sup e
0
Z 1
≤ kyk∞ max{1, eλ } e−λs ds.
0
Theorem 5.19. Let X be a Banach space, T ∈ L(X) and r(T ) its spectral
radius.
(i) r(T ) ≤ kT m k1/m ≤ kT k for all m ∈ N, in particular r(T ) = lim kT m k1/m .
m→∞
(ii) σ(T ) ⊆ {λ ∈ C : |λ| ≤ r(T )}.
(iii) If X is a complex Banach space, then there exists a λ ∈ σ(T ) such that
|λ| = r(T ), in particular
r(T ) = max{|λ| : λ ∈ σ(T )}.
where the series on the right hand side converges in norm. In particular, for
every ϕ ∈ L(X)0
∞
X
ϕ(λ − T )−1 = λ−(n+1) ϕ(T n ), |λ| > r(T ).
n=0
Hence λ 7→ ϕ(T − λ)−1 defines an analytic function for |λ| > r(T ). It follows
from complex analysis that then the equality in (5.3) holds for all λ in the largest
open ring
P∞ where λ 7→ ϕ(λ − T ) is analytic, that is for all λ > r(T ). In partic-
−(n+1)
ular, n=0 µ ϕ(T n ) converges for every ϕ ∈ L(X)0 , hence it is weakly
convergent, and therefore (µ−(n+1) ϕ(T n ))n∈N converges to 0. It follows that
(µ−(n+1) T n )n∈N is weakly convergent to 0, hence it is bounded (Corollary 3.9).
1 1 1
Let M ∈ R such that kµ−(n+1) kT n k < M , n ∈ N. Then kkT n k n < M n µ1+ n
1
for all n ∈ N, in particular r(T ) = lim kkT n k n ≤ µ.
n→∞
(iv) Recall that kT T ∗ k = kT k2 for a normal operator T (Theorem 4.39). Hence
Note that in general r(T ) < kT k, for example r(T ) = 0 for every nilpotent
linear operator.
(i) λ ∈ σp (T ) =⇒ λ ∈ σp (T 0 ) ∪ σr (T 0 ).
(ii) λ ∈ σr (T ) =⇒ λ ∈ σp (T 0 ).
Chapter 5. Spectrum of linear operators 107
(i) T is selfadjoint.
(ii) rg(λ − T ) = H for all z ∈ C \ R.
(iii) rg(±i − T ) = H.
(iv) There exist z± ∈ C with Im z+ > 0 and Im z− < 0 such that rg(z± − T ) =
H.
(v) σ(T ) ⊆ R.
(vi) T is closed and ker(±i − T ∗ ) = H.
Proof. (i) =⇒ (ii) Let λ ∈ C\R. Then rg(λ−T ) 6= H is closed by Theorem 5.22
and λ∗ ∈/ σp (T ). It follows by Theorem 4.73 that
(iv) =⇒ (v) Let z± ∈ C with Im z+ > 0 and Im z− < 0 such that rg(z± − T ) =
H. By Theorem 5.22, it follows that z± −T is injective and its inverse is bounded
by |=z± |. Hence, by Lemma 5.7, every λ ∈ C with |λ − z± | < |=z± | belongs to
ρ(T ). Given any λ ∈ C \ R, repeating the argument above finitely many times
shows that λ ∈ ρ(T ).
(v) =⇒ (ii) is obvious.
(vi) =⇒ (iii) Since T is closed, the range of ±i − T is closed by Theorem 5.22.
Therefore rg(±i − T ) = rg(±i − T )⊥⊥ = ker(∓i − T ∗ )⊥ = {0}⊥ = H.
(i) =⇒ (vi) Since T = T ∗ , it is closed and C \ R ⊆ ρ(T ), in particular ker(±i −
T ) = {0}.
Analogously, we find a characterisation of essentially selfadjoint operators.
kxn k = 1 and kR(λn , T )xn k ≥ 12 kR(λn , T )k. From Lemma 5.7 we know
that kR(λn , T )k ≥ dist(λn1,σ(T )) . Set yn := kR(λn , T )k−1 R(λn , T )xn . Then
yn ∈ D(T ) and kyn k = 1 for all n ∈ N. Moreover
k(λ − T )yn k ≤ k(λ − λn )yn k + k(λn − T )yn k
= |λ − λn | + kR(λn − T )xn k−1
≤ |λ − λn | + 2kR(λn − T )k−1 −→ 0, n → ∞.
Hence λ ∈ σap (T ).
(iii) By Theorem 5.23 the spectrum of a selfadjoint operator is real, so σ(T ) =
∂σ(T ) ⊆ σap (T ) ⊆ σ(T ).
Lemma 5.27. Let H be Hilbert space and T ∈ L(H) selfadjoint. Then σ(T ) ⊆
[m, M ] where m := inf{hT x , xi : kxk = 1} and M := sup{hT x , xi : kxk = 1}.
Moreover, m, M ∈ σ(T ).
Proof. Obviously, 0 ∈ K(X, Y ) and Remark 5.30 (ii) implies that the linear
combination of compact operators is compact. Now let (Tn )n∈N ⊆ K(X, Y ) a
Cauchy sequence. Since L(X, Y ) is complete, there exists a T ∈ L(X, Y ) such
that Tn → T . We have to show T ∈ K(X, Y ). Take an arbitrary bounded
sequence (xn )n∈N ⊆ X and choose M ∈ R such that kxn k ≤ M , n ∈ N.
(1) (1)
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
(k−1) (k) (n)
(xn ) such that (Tk xn )n∈N converges. Let (yn )n∈N = (xn )n∈N the diagonal
sequence. Then, for every k ∈ N, the sequence (Tk yn )n∈N converges. Let ε > 0.
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 yn − T ym k ≤ kT yn − Tk yn k + kTk yn − Tk ym k + kTk ym − T ym k
Mε ε Mε
≤ + + = ε.
3M 3 3M
Hence (T yn )n∈N is Cauchy sequence in the Banach space Y , hence convergent.
(i) A is bounded,
(ii) A is closed,
(iii) A is equicontinuous, that is,
Then A is compact.
fn : K → K, fn (y) := ϕn (y).
Then Tk is compact.
112 5.4. Compact operators
Proof. Obviously Tk is well-defined and bounded. Let (xn )n∈N ⊆ C([0, 1]) a
bounded sequence with bound C. Hence (Tk xn )n∈N is bounded. To show that
it is equicontinuous fix ε > 0. Since k is uniformly continuous, there exists
a δ > 0 such that |k(s, t) − k(s0 , t0 )| < ε if k(s, t) − (s0 , t0 )k < δ. Now for
t1 , t2 ∈ [0, 1] with |t1 − t2 | < δ and n ∈ N we obtain
Z 1
|Tk xn (t1 ) − Tk xn (t2 )| ≤ |k(s, t1 ) − k(s, t2 )||xn (s)| ds < εkxn k∞ ≤ Cε.
0
Proof. Let p := α(T ) and q := δ(T ). We divide the proof in several steps.
Step 1. rg(T p ) ∩ ker(T n ) = {0} for every n ∈ N0 .
To see this, choose x ∈ rg(T p ) ∩ ker(T n ). Then there exists a y ∈ X such that
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.
It follows that x = T p y = 0.
Step 2. X = rg(T n ) + ker(T q ) for every n ∈ N0 .
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 =
T n y + (x − T n y) ∈ rg(T n ) + ker(T q ).
Step 3. α(T ) ≤ δ(T ) = q.
Let x ∈ ker T q+1 . We have to show x ∈ ker T q . By step 2, with n = p,
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 ).
(ii) The spectrum of T is at most countable and 0 is the only possible accu-
mulation point.
Hence
1
kT yn − T ym k = kλn yn −(λn − T )yn − T ym k ≥ . (5.5)
| {z } 2
∈Un−1
116 5.4. Compact operators
Therefore (T xn )n∈N does not contain a convergent which contradicts the as-
sumption that T is compact.
(iii) and (iv) follow from Theorem 5.42.
(v) By Schauder’s theorem T 0 is compact (theorem 5.33) Hence for λ ∈ C it
follows that
(i) For every y ∈ X the equation (λ−T )x = y has exactly one solution x ∈ X.
(ii) (λ − T )x = 0 has a non-trivial solution x ∈ X.
(A) (λ − T )x = y, (C) (λ − T 0 )ϕ = η,
(B) (λ − T )x = 0, (D) (λ − T 0 )ϕ = 0.
Then
(a) For all y ∈ X and η ∈ X 0 the equations (A) and (C) have exactly one
solution (in particular (B) and (D) have only the trivial solutions).
Chapter 5. Spectrum of linear operators 117
(b) (B) and (D) have non-trivial solutions. In this case dim(ker(λ−T )) =
dim(ker(λ − T 0 )) > 0 and (A) and (C) have solutions if and only if
ϕ(y) = 0 for all solutions ϕ of (D),
η(x) = 0 for all solutions x of (B).
The λn can be chosen such that |λ1 | ≥ |λ2 | ≥ · · · > 0. The only possible
accumulation point of the sequence (λn )n∈N is 0.
(ii) If P0 is the orthogonal projection on ker T , then
N
X
x = P0 x + hx , en i en , x ∈ H. (5.7)
n=1
(iii) If λ ∈ ρ(T ), λ 6= 0
N
−1 −1
X hx , en i
(λ − T ) x=λ P0 x + en , x ∈ H.
λ −λ
n=1 n
118 5.4. Compact operators
It follows that
X
kT x − T hx , en i en k = kTn+1 xn+1 k ≤ |λn+1 |kxk −→ 0, n → ∞.
en ∈B1 ∪...Bn
Proof. If the series is a finite sum, the assertion is clear. Now assume
P∞ that the
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
k → ∞ the claim follows from
Xk
T − λn Pn
= sup{|λn | : n ≥ k + 1} = |λk+1 |.
n=1
Chapter 5. Spectrum of linear operators 119
Note that the theorem does not imply that there cannot be non-compact oper-
ators A ∈ L(H) such that A2 = T . In Corollary 5.60 we will show that every
bounded positive selfadjoint operator has a unique positive root.
Proof of Theorem 5.49. Recall that a linear operator T is positive if and only if
hT x , xi ≥ 0 for all x ∈ H. Let P0 , λn and en as in (5.7). Then (i) follows from
DX X E X
hT x , xi = λn hx , en i en , P0 x + λn hx, en , ein = λn |hx , en i|2 ≥ 0.
n n n
P 1/k k
For the proof of (ii) define R = n λn h· , en i en . Obviously R = T . To
show uniqueness, assume that there exists a compact selfadjoint positive linear
k
P S such that S = T . Since S is compact, it has a representation
operator
S = n µn Qn with pairwise orthogonal projections Qn . By assumption
X
T = Sk = µkn Qn .
n
Note that |T | can be defined more generally for positive selfadjoint operators
on a Hilbert space H, see Definition 5.61.
A representation similar to (5.6) exists for arbitrary compact operators.
Proof. (i) Let (ϕn )n∈N ⊆ H1 a ONS such that, see Theorem 5.47,
N
X N
X
|T |x = sn hx , ϕn iϕn , T ∗T x = s2n hx , ϕn iϕn .
n=1 n=1
1
Let ψn := xn T ϕn . Then (ψn )n∈N is an ONS in H2 because
1 1 1
hψn , ψm i = hT ϕn , T ϕm i = 2 hT ∗ T ϕn , ϕm i = 2 s2n δnm = δnm .
s2n sn sn
Moreover
1 s2
T T ∗ ψn = T T ∗ T ϕn = n T ϕn = s2n ψn .
sn sn
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}.
(ii) . . .
Proof. . . .
Chapter 5. Spectrum of linear operators 121
Theorem 5.55 (i) shows that for K ∈ HS(H1 , H2 ) the Hilbert-Schmidt norm
X 12
kKkHS := kK eα k2 for an ONB (eα )α∈A .
α∈A
is well-defined.
Proof of Theorem 5.55. (i)P Let K be a Hilbert-Schmidt operator and (eλ )λ∈Λ
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 ∗
is also a Hilbert-Schmidt operator.
2
X X
X ∗
XX
∗
kK ψβ k =2
hK ψβ , eλ i eλ
= |hK ∗ ψβ , eλ i|2
β∈B β∈B λ∈Λ λ∈Λ β∈B
XX X
2
= |hψβ , K eλ i| = kKψλ k2 < ∞.
λ∈Λ β∈B λ∈Λ
122 5.5. Hilbert-Schmidt operators
(ii) Let (eλ )λ∈Λ an ONS of H1 and (en )n∈N a subset containing all eλ with
K eλ 6= 0 (this family is at most countable by Lemma 4.27). For n ∈ N let Pn be
the orthogonal projection on {e1 , . . . , en }. Note that all Pn are compact because
they have finite-dimensional range. Since K is a Hilbert-Schmidt operator, we
find that
∞
X
kK − KPn k2 = kK(id −Pn )k2 ≤ kK(id −Pn )k2HS = kK en k2 −→ ∞,
m=n+1
X N
X N
X
kKϕλ k2 = kKϕn k2 ≤ kKk2HS = s2n < ∞,
λ∈Λ n=1 n=1
Lemma 5.56. The finite-rank operators are dense in the Hilbert-Schmidt op-
erators.
Theorem 5.57. Let H = L2 (0, 1) and T ∈ L(H). Then the following is equiv-
alent:
Chapter 5. Spectrum of linear operators 123
Proof. (ii) =⇒ (i) Let (en )n be an ONB of L2 (0, 1). Then also (en )n is an
ONB of L2 (0, 1) (where en denotes the to en complex conjugated function) and
we find
X∞ X∞ Z 1 Z 1 2 ∞ Z 1
X
2
|hk(·, t) , en i|2 dt
kT en k =
k(s, t) en (s) ds dt =
n=1 n=1 0 0 n=1 0
Z 1X∞
= |hk(·, t) , en i|2 dt (5.9)
0 n=1
Z 1
= kk(·, t)k2 dt (5.10)
0
Z 1 Z 1
= |k(s, t)|2 ds dt = kkkL2 (0,1)2 .
0 0
We will show that the range of Ψ is dense in HS(H). By Lemma 5.56 it suffices
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
n0
X n0 Z 1
X Z 1X n0
T f (t) = hf , xn iyn (t) = f (s)xn (s)yn (t) ds = xn (s)yn (t) f (s) ds.
n=1 n=1 0 0 n=1
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
sequence in H, hence its limit exists. Using the continuity of Ψ we find
S = lim Sn = lim ΨΨ−1 Sn = Ψ lim Ψ−1 Sn ∈ rg(Ψ).
n→∞ n→∞ n→∞
124 5.6. Polar decomposition
Proof. Note that (a+b)2 = a2 +b2 +2ab = a2 +b2 −(a−b)2 +a2 +b2 ≤ 2(a2 +b2 )
for a, b ∈ R.
(i) Let S, T ∈ HS(H1 , H2 ) and λ ∈ C. Then obviously λS ∈ HS(H1 , H2 ). To
show that S + T ∈ HS(H1 , H2 ) fix an ONS (eλ )λ∈Λ of H1 . Using the above
remark is follows that
X X X
k(S + T ) eλ k2 ≤ (kS eλ k + kT eλ k)2 ≤ 2 kS eλ k2 + kT eλ k2 < ∞.
λ∈Λ λ∈Λ λ∈Λ
id −A = T = R2 = (1 − X)2 = id −2X + X 2 .
1
X= (A + X 2 ). (5.11)
2
Step 1. Construction of a solution of (5.11).
We define
1 2
X0 := id, Xn := (A + Xn−1 ), n ∈ N.
2
Note that every Xn is a polynomial in A with positive coefficients and that
Xn Xm = Xm Xn for all n, m ∈ N. Since A is positive, this implies that all Xn
are positive. We will show the following properties of the sequence (Xn )n∈N by
induction.
(ii) kXn k ≤ 1.
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
1 1 1
Xn+1 − Xn = (A + Xn2 ) − (A + Xn−1
2
) = (Xn2 − Xn−1
2
)
2 2 2
1
= (Xn − Xn−1 )(Xn + Xn−1 ).
2
Since by induction hypothesis both terms in the second line are polynomials in
A with positive coefficients, (i) is proved for n + 1. (ii) follows from kXn+1 k ≤
1
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→∞
1 (see Exercise 4.25).
Now let S ∈ L(H) with ST = T S. By definition of A, then also SA = AS and
Xn S = Xn S for all Xn since the Xn are polynomials in A. For every x ∈ H we
therefore obtain
1 1 1
X = s- lim Xn = s- lim (A + Xn2 ) = (A + s- lim Xn2 ) = (A + X 2 ).
n→∞ n → ∞2 2 n→∞ 2
Setting R = id −X we obtain a bounded selfadjoint solution of R2 = T with
0 ≤ R ≤ id.
Step 2. Uniqueness of the solution.
Let R0 ∈ L(H) be solution of R2 = T with R0 ≥ 0. Then R and R0 commute
because
It follows that
Since both operators on the left hand side are non-negative, it follows that both
of them are 0 and therefore
(R − R0 )4 = (R − R0 )R(R − R0 ) − (R − R0 )R0 (R − R0 ) = 0.
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 .
is unitary.
Chapter 5. Spectrum of linear operators 127
We define
Appendix A
Lp spaces
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
the pointwise limit exists and is χQ∩[0,1] which is not Riemann integrabel. This
example shows that k·k1 is only a seminorm on R([a, b]), that Cauchy sequences
do not need to converge and that in general pointwise limit and integral cannot
be exchanged. The pointwise limit of a sequence of Riemann integrable functions
does not need to be Riemann integrable.
Recall that the Riemann integral of a function f : [a, b] → R is obtained as
the limit of Riemann sums when the interval [a, b] is divided in small pieces.
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
pre-images of intervals can be measured in some sense.
(a) ∅ ∈ Σ,
(b) A ∈ Σ =⇒ T \ A ∈ Σ,
(c) A, B ∈ Σ =⇒ A ∪ B ∈ Σ.
S
(iv) Σ ⊂ PT is called a σ-algebra if it is a algebra and n∈N An ∈ Σ for all
(An )n∈N ⊆ Σ.
(i) µ(∅) = 0,
S∞ P∞
(ii) (An )n∈N ⊆ A with pairwise disjoint An =⇒ µ( n=1 An ) = n=1 µ(An ).
Obviously, the intersection of rings is again a ring and PT is a ring. Hence, given
a family U of subsets of T , there exists a smallest ring containing U, namely the
intersection of all rings that contain U. This ring is called the ring generated by
U. Analogously the σ-ring, the algebra and the σ-algebra generated by U are
obtained.
Chapter A. Lp spaces 131
The aim is to assign a measure µ(U ) to every Borel set U ⊆ R such that the
measure of intervals is its length.
µ : Σ → [0, ∞]
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
is a pre-measure on A.
Proof. Obviously A is a ring, for every A ∈ A the characteristic function χA is
S integrable and µ(∅) = 0. S
Riemann Now let (An )Pn∈N ⊆ A be pairwise disjoint
n n
with n∈N An ∈ A. Obviously, µ( n=1 An ) = n=1 µ(An ) for every n ∈ N.
For n ∈ N define Bn := A \ (A1 ∪ · · · ∪ An ). Obviously, Bn ∈ A and
n
[
µ(A) = µ( Ak ) + µ(Bn ).
k=1
S∞
To prove that µ(A) = µ( k=1 Ak ) it suffices to show that lim µ(Bn ) = 0. Fix
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 .
By construction, B1 ⊇ B2 ⊇ B3 . . . , hence
n
[ n
[
µ(Bn \ Dn ) = µ(Bn \ (C1 ∩ · · · ∩ Cn ) = µ( (Bn \ Ck )) ≤ µ( (Bk \ Ck ))
k=1 k=1
n
X n
X
≤ µ(Bk \ Ck ) < 2−k ε = ε.
k=1 k=1
T∞ T∞
On the other hand, n=1 Dn ⊆ n=1 Bn = ∅. Since all Dn are compact and
D1 ⊇ D2 ⊇ . . . , there exists an K ∈ N such that Dn = ∅, n ≥ K. Hence
µ(Bn ) = µ(Bn \ Dn ) < ε for all n ≥ k.
132 A.1. A reminder on measure theory
In order to measure all Borel sets, we have to show that the pre-measure of
Example A.5 can be extended to the Borel sets.
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
to the set of the µ∗ -measurable sets is a measure.
µ∗ (Z) = µ∗ (Z ∩ A) + µ∗ (Z \ A), Z ∈ A.
Proof. Properties (i) and (ii) of an outer measure are clear. Now let (An )n∈N ⊆
j
S∞ andj ε > 0. Then there exists a family (Bn )n,j∈N ⊆ A such that An ⊆
PT
B
j=1 n and
∞
X ε
µ(Bnj ) ≤ µ∗ (An ) + , n ∈ N.
j=1
2n
Bnj and
S S
By construction A := n∈N An ⊆ n,j∈N
∞
X X
µ∗ (A) ≤ µ(Bnj ) ≤ µ∗ (An ) + ε.
n,j=1 n=1
Chapter A. Lp spaces 133
P∞
Note that (Bnj )n,j∈N is countable, hence we have proved µ∗ (A) ≤ n=1 µ∗ (An ).
∗
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
∞
X
µ(An ) ≤ µ∗ (A) + ε. n ∈ N.
n=1
S
Since A = n∈N (An ∩ A) and µ is a pre-measure on A, it follows that
∞
X ∞
X
µ(A) ≤ µ(An ∩ A) ≤ µ(An ) ≤ µ∗ (A) + ε. n ∈ N.
n=1 n=1
Since this is true for all ε > 0, it follows that µ(A) ≤ µ∗ (A).
Proof. . . . . . .
Proof of Theorem A.6. It suffices to show that the set of the µ∗ -measurable sets
contains A . . .
The measure on the completion of the Borel sets in R is the Lebesgue measure,
usually denoted by λ. .
A.2 Integration
In the following, I is always an interval in R.
It is easy to see that simple functions are measurable. Note, however, that the
sum representation of a simple function is not unique.
The next theorem lists important properties of measurable functions.
The theorem says that the set of the measurable functions are a vector space
and that it is stable under taking pointwise limits.
Next we introduce the integral for positive functions.
Of course, it must be proved that the integral in (i) does not depend on the sum
representation of the simple function, and that the limit in (ii) does not depend
on the chosen sequence of simple functions.
Chapter A. Lp spaces 135
For Lebesgue integrals much stronger convergence theorems hold than for the
Riemann integral. The most important convergence theorems are the following.
is measurable and
Z Z
f dλ = lim fn dλ.
I n→∞ I
A.3 Lp spaces
In the following, Ω is always an open subset of Rn .
It is easy to see that L∞ (Ω) is a vector space. For 1 ≤ p < ∞ this follows from
Z Z Z
|f + g| dλ ≤ (|f | + |g|) dλ ≤ (2 max{|f |, |g|})p dλ
p p
Ω Ω Ω
Z Z
p
≤2 max{|f | , |g| } dλ ≤ 2p
p p
|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:
1 1
Theorem A.24. Let 1 ≤ p ≤ ∞ and q the conjugated exponent, i. e., p+q = 1.
If f ∈ Lp (Ω) and g ∈ Lq (Ω), then f g ∈ L1 (Ω) and
kf gk1 ≤ kf kp kgkq .
Note that Lp (Ω) is only a seminormed space, because there are non-zero func-
tions f with kf kp = 0.
Theorem A.27. Let 1 ≤ p < ∞ and Ω ∈ Rn open. Then the test functions
Appendix B
Exercises
(i)X is complete.
(ii)Every absolutely convergent series in X converges in X.
V + W := {v + w : v ∈ V, w ∈ W }
is a closed subspace of X.
Z b
(a) Qn (f ) → f (t) dt, n → ∞, para todo f ∈ C[a, b].
a
Z b
(b) Qn (p) → p(t) dt, n → ∞, para todo polinomio p : [a, b] → K y
Pn a (n)
sup k=1 |αk | < ∞.
n∈N
6. Sea X = `2 (N) y
9. Sea X un espacio normado. Una sucesión (xn )n∈N ⊆ X es una sucesión débil
de Cauchy si para todo ϕ ∈ X 0 la sucesión (ϕ(xn ))n∈N es una sucesión de
Cauchy en K.
(a) Sea x = (xn )n∈N una sucesión acotada en X. Muestre que x es una
sucesión débil de Cauchy si y solo si existe un subconjunto denso U 0 de
X 0 tal que (ϕ(xn ))n∈N es una sucesión de Cauchy para todo ϕ ∈ U 0 .
(b) Toda sucesión débil de Cauchy en X es acotada.
(a) Muestre que (X, k · k)0 = (X, σ(X, X 0 ))0 . Es decir: un funcional lineal
ϕ : X → K es continua con respecto a la topologı́a inducida por k · k si
y sólo si es continua con respecto a la topologı́a débil.
w
(b) Sean (xn )n∈N ⊆ X, x0 ∈ X y (ϕn )n∈N ⊆ X 0 , ϕ0 ∈ X 0 tal que xn −
→ x0
w∗
y ϕn −−→ ϕ0 . Muestre
Halle una condición necesaria y suficiente spobre w para que h· , ·iw sea
un producto interno. Bajo qué condición la norma inducida por h· , ·iw es
equivalente a la norma usual de L2 ?
6. ¿Existe algún producto interno h· , ·i en C[0, 1] tal que hx, xi = kxk2∞ para
todo x ∈ C[0, 1] ?
146
¿El converso es cierto en general? ¿Hay algún caso para el que se tenga?
(b) Si x 6= 0, y 6= 0 y x ⊥ y muestre que el conjunto {x, y} es linealmente
independiente.
¿Como se puede generalizar este resultado?
(c) x ⊥ y, si y solo si kx + αyk ≥ kxk para todo escalar α.
11. Muestre que W m (Ω), H m (Ω) y H0m (Ω) son espacios de Hilbert.
Para el problema 4.10: Para Ω ⊆ R definimos el conjunto de funciones de
prueba
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 .
B(x, y) = hT x , yi, x, y ∈ H.
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,
m
1 X
ym = x nk ,
m
k=1
converge.
18. Sea H un espacio de Hilbert separable, (xn )n∈N una base ortonormal de H,
y, (yn )n∈N una sucesión tal que:
∞
X
||xn − yn || < 1
n=1
(a) (hj )j∈N0 is a orthonormal system in L2 [0, 1] and (ψn,k )n,k∈Z is a or-
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-
tion on the subspace
V ⊆W ⇐⇒ PV = PV PW = PW PV .
23. Sea H un espacio de Hilbert separable, (xn )n∈N una base ortonormal de H,
y, (yn )n∈N una sucesión tal que:
∞
X
||xn − yn || < 1
n=1
25. Sea H un espacio de Hilbert y (Tn )n∈N una sucesión acotada y monótonamente
creciente de operadores autoadjuntos. Muestre que la sucesión converge en
el sentido fuerte a un operador autoadjunto.
26. Sea (Pn )n∈N una sucesión monótona de proyecciones ortogonales en un es-
pacio de Hilbert H. Muestre que (Pn )n∈N converge en el sentido fuerte a
una proyección ortogonal P y además
S
(a) rg P = n∈N rgPn si Pn es creciente.
T
(b) rg P = n∈N rgPn si Pn es decreciente.
(c) T ∗ es cerrado.
(d) Si T es clausurable, T ∗ es densamente definido y T ∗∗ = T .
[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.
[Kat95] Tosio Kato. Perturbation theory for linear operators. Classics in Math-
ematics. Springer-Verlag, Berlin, 1995. Reprint of the 1980 edition.
[Rud87] Walter Rudin. Real and complex analysis. McGraw-Hill Book Co., New
York, third edition, 1987.
153
154 Bibliography