Introduction To FuntionalAnalysis
Introduction To FuntionalAnalysis
Faculty of Science
Department of Mathematics
Lecture Notes
MEAS Len
©2022
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Contents
Preface 1
1 Foundations 3
1.1 Lebesgue Integral 3
1.2 Metric Space 27
1.3 Banach Space 46
1.4 Bounded Linear Operators 54
1.5 Hilbert Space 65
1.6 Dual Space 72
2 Main Theorems 77
2.1 Baire’s Category Theorem 77
2.2 Uniform Boundedness Principle 79
2.3 Open Mapping Theorem 84
2.4 Closed Graph Theorem 85
2.5 Hahn-Banach Theorem 87
3 Compact Operators 97
3.1 Compact operators 97
3.2 Spectral theorem for self-adjoint compact linear operators 104
References 106
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
ii
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Preface
The aim of this book is to build a foundation of functional analysis for undergraduate and the
beginning graduate students in science, engineering and mathematics who are interested in the
research in analysis.
The focus of the first part of this book is on the presentation of the fundamental concepts in
functional analysis: Banach space, linear operator, Hilbert space and dual space. The emphasis
in the second part is on the main results such as uniform boundedness principle, open mapping
theorem, closed graph theorem and Hahn-Banach theorem. The third part is devoted to the
concept of compact operators which play a crucial role in the theory of integral equations and
partial differential equations and in mathematical physics. The relation of compactness with
weak convergence and reflexivity, the spectral properties of self-adjoint compact linear oper-
ators are hightlighted. Specifically, the spectral theorem for self-adjoint comapct operator is
presented.
Without the wonders of LaTeX, this book would never have been prepared. I can only hope
that the errors are unintentional. For this reason, I am particularly interested in receiving cor-
rections, comments and suggestions via email: [email protected]
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Chapter 1
Foundations
This chapter will focus on the fundamental concepts in functional analysis: Banach space,
linear operator and dual space. The concepts of continuity, connectedness and compactness
will also be covered. In addition, the distinction features of finite and infinite dimensional
normed spaces will also be discussed.
Example 1.1
I = {(x1 , x2 ) ∈ R2 | 1 ≤ x1 ≤ 3, −2 ≤ x2 ≤ 4}
is a closed interval in R2 , while
J = {x ∈ R1 | 2 ≤ x ≤ 6}
is a closed interval in R1 .
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
The volume of a closed interval in Rn is v(I) = ni=1 (bi − ai ), where the constants ai and bi are
Q
all finite, and that ai < bi for all i.
More precisely, let S = {Ik } be a countable (finite or countably infinite) collection of closed
intervals in Rn . We say that S is a covering of A by closed intervals if A ⊆ ∪Ik . We set
σ(S ) = v(Ik ). If the series v(Ik ) diverges, set σ(S ) = +∞.
P P
Definition 1.1
By Definition 1, m∗ (A) is always greater or equal to 0. Also, it follows that if S is any covering
of A by closed intervals, then m∗ (A) ≤ σ(S ) ≤ +∞. In other words, for any set A ⊆ Rn , 0 ≤
m∗ (A) ≤ +∞.
Remark 1.1
An important feature of the definition of Lebesgue outer measure is that given any set A
and any given ϵ > 0, there is a covering S of A by closed intervals such that
σ(S ) ≤ m∗ (A) + ϵ.
Example 1.2
Example 1.3
The Lebesgue outer measure of ∅ is 0. To see this, let ϵ > 0 be given. Then S = {[−ϵ, ϵ]}
is a covering of ∅ by closed intervals. Therefore,
m∗ (∅) ≤ σ(S ) = 2ϵ
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
m∗ (A ∪ B) ≤ m∗ (A) + m∗ (B).
Propoistion 1.1
Example 1.4
Let A = [0, 1]. The Lebesgue outer measure of A is 1. (This is just the length of interval
[0, 1]).
Example 1.5
m∗ (B) ≤ 3.
Consequently, m∗ (B) = 3.
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
A set E ⊆ Rn is Lebesgue measurable if for every ϵ > 0 there is an open set G so that
E ⊆ G and
m∗ (G \ E) < ϵ.
In this case, we define the Lebesgue measure of E, denoted by m(E), to be
m(E) = m∗ (E).
Example 1.6
Example 1.7
If G is open set, then m∗ (G \ G) = m∗ (∅) = 0 < ϵ for every ϵ > 0. Consequently, every
open set in Rn is Lebesgue measurable.
Theorem 1.1
As a consequence:
One has any intervals are Lebesgue measurable and m(I) = m∗ (I) = v(I).
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
M
Let {In }n=1 be a finite collection of pairwise non-overlapping closed intervals. Then
M M
[ X
m In =
v(In ).
n=1 n=1
In general, every open set G ⊆ Rn can be written as a countably infinite union of non-overlapping
closed intervals G = ∞
S
k=1 Ik with
X∞
m(G) = v(Ik ).
k=1
n
What are Lebesgue measurable subsets of R ?
Lebesgue measurable subsets of Rn are
2. Open sets
3. Closed sets
4. Intervals
is Lebesgue measurable.
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Theorem 1.3
It says that the measure of the union of disjoint Lebesgue measurable sets is the sum of mea-
sures.
Remark 1.2
(i). X ∈ M.
(ii). A \ B ∈ M whenever A, B ∈ M.
∞
S
(iii). An ∈ M whenever (An )n≥1 ⊂ M.
n=1
Example 1.8
Example 1.9
Let X = R. Borel σ-algebra of R, B is the smallest σ-algebra which contains all the open
subsets of R.
Borel sets: (a, b), [a, b], {a}, (a, b], [a, b)
∞ !
\ 1
(a, b) = a, b +
n=1
n
∞ !
\ 1 1
{a} = a − ,a +
n=1
n n
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
∞ !
\ 1 1
[a, b] = a − ,b + .
n=1
n n
Definition 1.4
a µ(∅) = 0.
The triple ([a, b], L([a, b]), m) is a measure space m : L([a, b]) → [0, ∞) by
1. m(∅) = 0;
N
2. if A =
S
In disjoint uniform of intervals,
n=1
N N
[ X
m In =
l(In )
n=1 n=1
Example 1.11
Let (X, M, µ) be a measure space with µ(X) = 1 (probability space, µ: probability mea-
sure).
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Proof. (a). B can be decomposed as B = A ∩ (B \ A), a disjoint union. Thus the additivity of µ
yields
µ(B) = µ(A ∪ (B \ A)) = µ(A) + µ(B \ A) ≥ µ(A).
∞
X n
X
µ(A) = µ(B j ) = lim µ(B j ) = lim µ(An ).
n→∞ n→∞
j=1 j=1
10
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Moreover,
∞
[
µ(A1 \ A) = µ Cn = lim µ(Cn )
n→∞
n=1
= lim (µ(A1 ) − µ(An ))
n→∞
µ(A1 ) − µ(A) = µ(A1 ) − lim µ(An )
n→∞
(d). Let B1 = A1 , B2 = A1 ∪ A2 , · · · , Bn = A1 ∪ A2 ∪ · · · ∪ An .
∞
[ ∞
[
Bn = An .
n=1 n=1
Therefore,
∞
[
µ Bn = lim µ(Bn ) as (Bn )n≥1 is increasing
n→∞
n=1
n
[
= lim µ Ai
n→∞
i=1
Hence,
∞ n ∞
[ X X
µ An ≤ lim
µ(Ai ) = µ(An ).
n→∞
n=1 i=1 n=1
Measurable Function
Definition 1.5: Measurable function
Let (X, M, µ) be a measure space and let f : X → [−∞, ∞]. We say that f is a measurable
function (or M-measurable) if for every a ∈ R, the set {x : f (x) > a} is a measurable set
(i.e., element of σ-algebra).
Example 1.12
1. Any continuous function f is measurable with measure space (R, L, m) since the
inverse image of any open set is open and L contains all open subsets of R.
11
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Since each of the sets R, Q and ∅ is Lebesgue measurable, it follows that the Dirichlet
function is Lebesgue measurable.
Definition 1.6
Let (X, M, µ) be a measure space and let E ⊆ X. The characteristic function of E denoted
by 1E , defined by
1, x∈E
1E =
0, x < E.
12
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Theorem 1.5
Theorem 1.6
Let (X, M, µ) be a measure space and let f : X → [−∞, ∞]. The following are equivalent:
∞ n o
• (a) ⇔ (b) : {x : f (x) ≥ a} = x : f (x) > a − 1
T
Proof. k
∈ M.
k=1
∞ n o
• (b) ⇔ (a) : {x : f (x) > a} = x : f (x) ≥ a + 1
S
k
∈ M.
k=1
Propoistion 1.2
Let (X, M, µ) be a measure space and let f, g : X → R be measurable functions, and let
c ∈ R. Then | f |, c f, f + g, f g and f /g (g , 0) are measurable.
13
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Proof. We have
X
if a < 0
{x : | f (x)| > a} =
{x : f (x) > a} ∪ {x : f (x) < −a} if a ≥ 0
Since X ∈ M and f is measurable it follows that {x : f (x) > a} ∈ M and {x : f (x) < −a} ∈ M;
hence
{x : f (x) > a} ∪ {x : f (x) < −a} ∈ M.
If c = 0 it is trivial. Assume c , 0. We have
{x : f (x) < a/c} if c < 0
{x : c f (x) > a} =
{x : f (x) > a/c} if c > 0
Since f and g are measurable, the sets {x : f (x) > r} and {x : g(x) > a − r} are rational for
every rational r. Thus the intersection of these sets is measurable and the countable union is
also measurable. The function f 2 is measurable since
X
if a < 0
{x : f (x) > a} =
2
√ √
{x : f (x) < − a} ∪ {x : f (x) > a} if a ≥ 0
It follows that
1
f g = [( f + g)2 − f 2 − g2 ]
2
is measurable.
Finally, if g , 0
{x : g(x) < 1/a} if a > 0
{x : 1/g(x) > a} = {x : g(x) > 0} if a = 0
{x : g(x) > 0} ∪ {x : g(x) < 1/a} if a < 0
Propoistion 1.3
Let f, g be measurable functions. Then max{ f, g}, min{ f, g} are measurable functions.
14
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Remark 1.3
As a consequence, the positive part of f , f+ (x) = max{ f (x), 0} is measurable and the
negative part of f is f− = max{− f (x), 0} is measurable. Note that f = f+ − f− and | f | =
f+ + f− .
We define
sup fn (x) : = sup{ fn (x)}.
n∈N n∈N
inf fn (x) : = inf { fn (x)}.
n∈N n∈N
(lim sup fn )(x) := lim sup{ fk (x) : k ≥ n} = inf sup{ fk (x) : k ≥ n} .
n→∞ n≥1
(lim inf fn )(x) := lim inf{ fk (x) : k ≥ n} = sup inf{ fk (x) : k ≥ n} .
n→∞ n≥1
Theorem 1.7
Let ( fn )n∈N be a sequence of measurable functions. Then supn∈N fn , inf n∈N fn , lim sup fn and
lim inf fn are measurable.
so the supremum and infimum are measurable. Moreover, by definition of limsup and liminf,
we get that both are measurable. □
15
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Theorem 1.8
which is measurable. □
Lebesgue Integral
Theorem 1.9: Fundamental Approximation
Let (X, M, µ) be a measure space and let f : X → [0, ∞] be measurable. Then there exists
a sequence of measurable, nonnegative, simple functions (sn ) with
Corollary 1.1
16
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Remark 1.4
Convention: 0 · ∞ = 0.
Example 1.13
if − 1 < x < 2
1,
if 2 ≤ x < 4
2,
φ(x) =
3, if 4 ≤ x ≤ 8
0,
otherwise.
Z
φ(x)dx = 1 · µ((−1, 2)) + 2 · µ([2, 4)) + 3 · µ([4, 8])
R
= 1 × 3 + 2 × 2 + 3 × 4 = 19.
Example 1.14
That is,
2, if 0 ≤ x ≤ 2
φ(x) = if 2 < x ≤ 3
3,
0,
otherwise
17
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Then
Z
φ(x)dx = 2.µ([0, 2]) + 3.µ((2, 3])
R
= 2(2 − 0) + 3(3 − 2)
= 7.
Example 1.15
That is,
2, if − 1 ≤ x ≤ 1
if 3 < x < 7
3,
φ(x) =
−1, if − 4 ≤ x < −3
0,
otherwise
Then
Z
φ(x)dx = 2.µ([−1, 1]) + 3.µ((3, 7)) − 1.µ([−4, −3))
R
= 2.(1 − (−1)) + 3.(7 − 3) + (−1).((−3) − (−4))
= 15.
Example 1.16
= cµ([a, b])
= c(b − a).
Example 1.17
18
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Remark 1.5
Remark 1.6
It follows that
µ(Qc ∩ [a, b]) > 0.
Example 1.18
19
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Property 1.3
Theorem 1.10
Let (X, M, µ) be a measure space and let s be a measurable, nonnegative, simple function.
The set function φ : M → [0, ∞] defined by
Z
φ(E) = sdµ
E
is a positive measure on M.
Then
Z n
X
φ(A) = sdµ = c j µ(C j ∩ A)
A j=1
n
∞ n ∞
X [ X X
= c j µ (C j ∩ Ak ) = µ(C j ∩ Ak )
cj
j=1 k=1 j=1 k=1
∞
XX n ∞
XZ ∞
X
= c j µ(C j ∩ Ak ) = sdµ = φ(Ak ).
k=1 j=1 k=1 Ak k=1
□
20
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Definition 1.9
Property 1.4
Remark 1.7
21
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
| f (x)|− f (x)
2
is the negative part of f . If E is a measurable set then we define
Z Z Z Z Z
f dµ = +
f dµ − f dµ =
−
f .1E dµ −
+
f − .1E dµ.
E E E X X
Propoistion 1.4
Let f and g be two functions that are integrable with respect to µ. If f (x) = g(x) a.e., then
Z Z
f dµ = gdµ.
X X
Proof. We first prove this in the case where both f and g are nonnegative. Let Z = {x ∈
X| f (x) , g(x)}. Then µ(Z) = 0. Let φ be a simple function with 0 ≤ φ ≤ f , say
n
ai 1Ei (x),
X
φ(x) =
i=1
where ai ≥ 0. Set
n
ai 1Ei \Z (x).
X
ψ(x) =
i=1
Then 0 ≤ ψ ≤ g and
Z n
X n
X Z
ψdµ = ai µ(Ei \ Z) = ai µ(Ei ) = φdµ.
X i=1 i=1 X
Thus, Z Z
φdµ ≤ gdµ.
X X
Taking the supremum over all simple functions 0 ≤ φ ≤ f,
Z Z
f dµ ≤ gdµ.
X X
The reverse inequality follows the same manner. Therefore, in this case
Z Z
f dµ = gdµ.
X X
Finally, the general result follows by considering the positive and negative parts of f and g. □
22
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Property 1.5
Proof. Note that the limit function f is measurable. The monotonicity of the integrals yields
Z Z Z
fn dµ ≤ fn+1 dµ ≤ f dµ
X X X
for every n. By monotone convergence of sequences, the limit of the integrals exists as a value
in [0, ∞], call it α. We have that Z
α≤ f dµ.
X
If α = ∞, the equality holds. Now assume that α < ∞. Consider a simple function s with
0 ≤ s(x) ≤ f (x) on X and fix 0 < δ < 1. For each n, define a set
• X= ∞
S
n=1 E n . Why?
23
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
S∞
(i) n=1 En ⊆ X
(ii) Now we prove that X ⊆ ∞ n=1 E n . If x ∈ X and f (x) = 0, then s(x) = 0 and
S
x ∈ En for all n. If 0 < f (x) < ∞, then δs(x) ≤ δ f (x) < fn (x) for all large n. If
f (x) = ∞, fn (x) → ∞ so δs(x) < fn (x) for all large n.
Proof. Define a sequence of functions (gn ) by gn (x) = inf k≥n { fk (x)}. Note that the functions
in this sequence is measurable and increasing by construction. The function lim inf fn is mea-
surable and we can write lim inf fn = limn→∞ gn . Applying Lebesgue’s monotone convergence
theorem 11, Z Z Z
lim inf fn dµ = lim gn dµ = lim gn dµ.
X X n→∞ n→∞ X
24
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
By the properties of infimum it follows that gn (x) ≤ fk (x) for every x ∈ X and all k ≥ n, from
which the monotonicity of the integral implies
Z Z
gn dµ ≤ inf fk dµ.
X k≥n X
Hence Z Z ! Z
lim inf fn dµ ≤ lim inf fk dµ = lim inf fn dµ.
X n→∞ k≥n X X
Example 1.19
For n = 1, 2, 3, . . . define
n if 0 < x ≤
1
fn (x) =
n
0 otherwise
Suppose that fn : X → [−∞, ∞] is a sequence of measurable functions such that the limit
function lim fn (x) = f (x) exists for every x ∈ X.
n→∞
Further suppose there exists a Lebesgue integrable function g such that | fn (x)| ≤ g(x) for
every x ∈ X. Then
fn dµ =
R R
(c). limn→∞ X X
f dµ.
25
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Proof. (a). Note that the function f is measurable. Since ( fn ) converges pointwise to f and
| fn (x)| ≤ g(x) on X, it must also be the case that | f (x)| ≤ g(x) on X. Then we have
Z Z
| f |dµ ≤ gdµ
X X
and f is integrable.
Hence Z
lim sup | fn − f |dµ ≤ 0.
X
However, | fn − f | ≥ 0 and thus
Z
lim inf | fn − f |dµ ≥ 0.
X
Therefore, we get
Z Z
0 ≤ lim inf | fn − f |dµ ≤ lim sup | fn − f |dµ ≤ 0
X X
26
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Remark 1.8
• d : X × X → R is a “distance function”.
Consequences:
Example 1.20
27
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Example 1.21
For any interger n > 1 and let 1 ≤ p ≤ ∞. The space (Rn , d p ) or (Cn , d p ) is a metric space,
where 1/p
n
if 1 ≤ p < ∞
P p
|xi − yi |
d p (x, y) =
i=1
max1≤i≤n |xi − yi | if p = ∞,
Remark 1.9
The trianlge inequality for the distance function is called Minkowski’s inequality. It is
straightforward to verify if p = 1 or p = ∞, but it is not obvoius if 1 < p < ∞.
The proof of the triangle inequality uses the following sequence of inequalities.
Lemma 1.1
xλ y1−λ ≤ λx + (1 − λ)y.
Let p ∈ (1, ∞) and define q ∈ (1, ∞) by 1p + 1q = 1. Then for any real numbers α, β > 0, it
holds
α p βq
αβ ≤ + .
p q
28
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Note that when p = q = 2 this is the Cauchy-Schwarz inequality. Note also that it also holds
when p = 1 and q = ∞.
Apply Hölder’s inequality for each of the two sums on the right, we get
Xn n
X n
1/p X 1/q
|xi ||xi + yi | ≤
p−1
|xi | p
|xi + yi |(p−1)q
29
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
and
n
X n
X n
1/p X 1/q
|yi ||xi + yi | p−1
≤ |yi | p
|xi + yi | p
.
i=1 i=1 i=1
Thus,
n
X n
X 1/p n
X n
1/p X 1/q
|xi + yi | ≤
p
|xi | p
+ |yi | p
|xi + yi | p
.
i=1 i=1 i=1 i=1
P 1/q
If x + y , 0, we divide both sides by n
i=1 |xi + yi | p to obtain
n
X 1/p n
X 1/p n
X 1/p
|xi + yi | p
≤ |xi | p
+ |yi | p
.
i=1 i=1 i=1
since 1 − 1
q
= 1
p
and note that if x + y = 0 the desired inequality is trivial. □
Example 1.22
Let 1 ≤ p ≤ ∞. The set l p of all real or complex sequences (xk )k≥1 satisfying
∞
X
|xk | p < ∞ (1 ≤ p < ∞),
k=1
sup |xk | < ∞ (p = ∞)
k∈N
30
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Example 1.23
and
L∞ (X) = f : f is measurable and ess sup | f | < ∞ .
Note that the function in L p is defined a.e. In other words, for a measurable function f , the set
of all functions which are a.e. equal to f can be identified as
31
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Let 1 ≤ p, q ≤ ∞ satisfying 1
p
+ 1
q
= 1. Then for f ∈ L p and g ∈ Lq it holds f g ∈ L1 and
Z Z1/p Z 1/q
| f g|dµ ≤ | f | dµ qp
|g| dµ (1 < p < ∞),
ZX X Z X
| f g|dµ ≤ ess sup | f | |g|dµ.
X X
Proof. We will consider the case where 1 < p, q < ∞ as the case p = 1 or p = ∞ is easily ver-
R 1/p R 1/q
p
ified. We can assume that X | f | dµ > 0 and X |g| dµ
q
> 0. Using Young’s inequality
5 with
|f| |g|
α= 1/p and β = R 1/q
R
| f | p dµ |g|q dµ
X X
Z Z 1/2 Z 1/2
| f g|dµ ≤ 2
| f | dµ |g|2 dµ .
X X X
and for p = ∞,
ess sup | f + g| ≤ ess sup | f | + ess sup |g|.
32
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Apply Hölder’s inequality for each of the two integrals on the right, we get
Z Z 1/p Z 1/q
| f || f + g| dµ ≤
p−1 p
| f | dµ | f + g|(p−1)q
dµ
X X X
Z 1/p Z 1/q
= p
| f | dµ | f + g| dµ
p
X X
R 1/q
X
| f + g| dµ
p
= 0 the result is obvious.
If p = ∞, we have | f | ≤ ess sup | f | a.e. in X and |g| ≤ ess sup |g| a.e. in X . Therefore,
| f + g| ≤ ess sup | f | + ess sup |g| a.e. in X, so that
A sequence (xn ) in a metric space (X, d) converges to x ∈ X if for every ε > 0, there exists
N ∈ N such that
d(xn , x) < ε, for all n ≥ N.
Notation: lim xn = x or xn → x.
n→∞
Definition 1.13
A sequence (xn ) in (X, d) is a Cauchy sequence if, for every ε > 0, there exists N ∈ N such
that
d(xm , xn ) < ε, for all m, n ≥ N.
33
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Theorem 1.14
Suppose that (xn ) is a convergent sequence in a metric space (X, d). Then
Proof. 1. Suppose lim xn = l and lim xn = m. Let ε > 0. Then for n large enough
n→∞ n→∞
ε ε
d(l, m) ≤ d(l, xn ) + d(xn , m) < + = ε.
2 2
Thus 0 ≤ d(l, m) ≤ ε holds for any ε > 0. It follows that d(l, m) = 0, i.e, l = m.
2. Let ε > 0 be given. Since xn → l, there exists an Nε ∈ N such that xn ∈ Bε (l) for all
n ≥ Nε . If k ≥ Nε then nk ≥ k, so that nk ≥ Nε , and thus xnk ∈ Bε (l). This implies, by
definition, that xnk → l as k → ∞.
3. Let ε > 0 , since xn → l, ∃N such that ∀n ≥ N, d(xn , l) < ε/2. Then if m, n > N,
ε ε
d(xn , xm ) ≤ d(xn , l) + d(l, xm ) ≤ + = ε.
2 2
Thus (xn ) is a Cauchy sequence.
□
Propoistion 1.11
∃N1 , n ≥ N1 ⇒ xn ∈ Br (x)
∃N2 , n ≥ N2 ⇒ xn ∈ Br (y).
For n ≥ max(N1 , N2 ) this would result in xn ∈ Br (x) ∩ Br (y) = ∅, a contradiction. □
34
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Propoistion 1.12
Proof. If x , y then d(x, y) > 0. Let r := d(x, y)/2, then Br (x) is disjoint from Br (y) else we get
a contradiction: z ∈ Br (x) ∩ Br (y). Then
d(x, z) < r and d(y, z) < r.
Hence
d(x, y) ≤ d(x, z) + d(y, z) < 2r = d(x, y).
□
Example 1.24
Definition 1.14
xn → x in X ⇒ f (xn ) → f (x) in Y.
Remark 1.10
Theorem 1.15
(i). f is continuous
35
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Proof. (i) ⇒ (ii): Suppose (ii) is false, then there is a point x ∈ X and an ε > 0 such that
arbitrarily small changes to x can lead to large variations in f (x), ∀δ > 0, ∃x′ , dX (x, x′ ) < δ and
dY ( f (x), f (x′ )) ≥ ε.
In particular, let δ = n1 , there is a sequence xn ∈ X satisfying dX (x, xn ) < n1 but dY ( f (x), f (xn )) ≥
ε. This means that xn → x, but f (xn ) ↛ f (x), contradicting statement (i).
(ii) ⇒ (iii): Note that (ii) can be written as
∀ε > 0, ∃δ > 0, ∀x ∈ X, x′ ∈ Bδ (x) ⇒ f (x′ ) ∈ Bε ( f (x))
or even as
∀ε > 0, ∃δ > 0, ∀x ∈ X, f (Bδ (x)) ⊆ Bε ( f (x)).
Let V be an open set in Y. To show that U := f −1 (V) is open in X, let x be any point of U, then
f (x) ∈ V, which is open. Hence,
f (x) ∈ Bε ( f (x)) ⊆ V,
and so
∃δ > 0, f (Bδ (x)) ⊆ Bε ( f (x)) ⊆ V.
In other words, x is an interior point of U, ∃δ > 0, Bδ (x) ⊆ f −1 (V) ⊆ U.
(iii) ⇒ (i): Let (xn ) be a sequence converging to x. Consider any open neighborhood Bε ( f (x))
of f (x). Then f −1 (Bε ( f (x)) contains x and is open set by (iii), so
∃δ > 0, x ∈ Bδ (x) ⊆ f −1 (Bε ( f (x)) ⇒ ∃δ > 0, f (Bδ (x)) ⊆ Bε ( f (x)).
But eventually all the points xn are inside Bδ (x),
∃N > 0, n > N ⇒ xn ∈ Bδ (x)
⇒ f (xn ) ∈ f (Bδ (x)) ⊆ Bε ( f (x))
⇒ dY ( f (xn ), f (x)) < ε.
This shows that f (xn ) → f (x) as n → ∞. □
Remark 1.11
36
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
When X and Y are homeomorphic, they are not only the same as a sets (bijection) but
also with repsect to convergence
xn → x ⇐⇒ f (xn ) → f (x),
and
A is open in X ⇐⇒ f (A) is open in Y.
The most vivid picture is that of “deforming” one space continuously and reversibly form
the other. The classic example is that a ‘teacup’ is homeomorphic to a “doughnut”.
Theorem 1.16
The proof of this theorem follows the Cauchy criterion in real analysis which states that a
sequence converges if and only if it is Cauchy.
Propoistion 1.13
√
Proof. Consider a sequence rationals xn converging to 2 in (R, d), specifically, we may as-
sume
√ 1
xn − 2 ≤
n
√
(This is possible since there are rationals number arbitrarily close to 2). Indeed, we can use
recurrence x1 = 1 and
xn−1 1
xn = + ,
2 xn−1
√
for n > 1. As (xn ) converges, then it is Cauchy. But its limits 2 < Q. □
Propoistion 1.14
37
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Proof. Let F ⊆ X be complete, i.e. any Cauchy sequence in F converges to a limit in F. Let
x ∈ F̄, with a sequence xn → x, xn ∈ F. Since convergent sequences are Cauchy and F is
complete, x must be in F. Thus F = F̄ is closed.
Conversely, let F be a closed in X and let (xn ) be a Cauchy sequence in F. Then (xn ) is a Cauchy
sequence in X, which is complete. Therefore, xn → x for some x ∈ X. In fact x ∈ F̄ = F. Thus
any Cauchy sequence of F converges in F. □
Remark 1.12
2. Two metric spaces may be homeomorphic yet one space be complete and the other
not. For example: R is homeomorphic to ]0, 1[. But ]0, 1[ is not closed in R, hence
not complete.
Note that a continuous function need not preserve completeness, or even Cauchy sequences; so
we need to strengthen continuity.
Definition 1.16
Remark 1.13
Propoistion 1.15
38
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Example 1.25
Lipschitz maps are uniformly continuous: since for any ε > 0, we can let δ = 2cε inde-
pendent of x to obtain d(x, x′ ) < δ ⇒ d( f (x),√f (x′ )) < cδ < ε. Not every uniformly
continuous function is Lipschitz. √For example,
√ x on [0, 1] is uniformly
√ continuous. If it
were Lipschitz, it would satisfy | x − 0| ≤ c|x − 0| which leads to x ≥ 1c .
Definition 1.17
Proof. Consider the iteration xn+1 := f (xn ) starting with any x0 in X. Note that
d(xn+1 , xn ) = d( f (xn ), f (xn−1 )) ≤ cd(xn , xn−1 ).
Hence, by induction on n,
d(xn+1 , xn ) ≤ cn d(x1 , x0 ).
So (xn ) is Cauchy since c < 1. As X is complete, xn converges x and by continuity of f ,
f (x) = f lim xn = lim f (xn ) = lim xn+1 = x.
n→∞ n→∞ n→∞
cn
Moreover, the rate of convergence is given at least by d(x, xn ) ≤ 1−c
d(x1 , x0 ).
Suppose there are two fixed points x = f (x) and y = f (y), then
d(x, y) = d( f (x), f (y)) ≤ cd(x, y)
implying d(x, y) = 0 since c < 1. □
39
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Remark 1.14
The assumption complete metric space is important. Consider (0, 1), |.| is not complete
metric space (the sequence (1/n)n≥1 ⊂ (0, 1) and converges to 0, making it a Cauchy se-
quence in (0, 1) ⊂ R, but 0 < (0, 1)). Consider the function f : (0, 1) → (0, 1), f (x) = 21 x.
Then for all x, y ∈ (0, 1)
1 1 1 1
| f (x) − f (y)| = x − y = |x − y| ≤ |x − y|
2 2 2 2
so f is a contraction with c = 12 . The fixed point equation for f , i.e., f (x) = x has no
solutions in (0, 1) since 21 x = x if and only if x = 0 < (0, 1).
A subset C of a metric space is disconnected when it can be divided into (at least) two
disjoint non-empty subsets C = A ∪ B such that each subset is covered exclusively by an
open set, i.e.,
A ⊆ U, B ∩ U = ∅, U open
B ⊆ V, A ∩ V = ∅, V open.
Theorem 1.18
Propoistion 1.16
Continuous functions map connected sets to connected sets. That is, f : X → Y continu-
ous, C ⊆ X connected, then f (C) is connected.
Proof. Let C be a subset of X, and suppose f (C) is disconnected into the non-empty disjoints
sets A and B convered exclusively by open sets U and V; that is,
f (C) = A ∪ B ⊆ U ∪ V, U ∩ B = ∅ = V ∩ A.
Then
C = f −1 (A) ∪ f −1 (B) ⊆ f −1 (U) ∪ f −1 (V),
40
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Theorem 1.19
[ iN
[
K⊆ Ai ⇒ ∃i1 , · · · , iN , K = Ai
i i=i1
Remark 1.15
In metric space,
Ai = Bεi (ai ).
Propoistion 1.17
41
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Propoistion 1.18
Proof. (i). Let F be a closed subset of a compact set K, and let the open set Ai cover F, then
[
K ⊆ F ∪ (X \ F) ⊆ Ai ∪ (X \ F).
i
But K is compact and therefore a finite number of these open sets are enough to cover it
N
[ N
[
K⊆ Ai ∪ (X \ F), so F = Ai .
i=1 i=1
(ii). Let the open sets Ai cover the finite union of compact sets K1 ∪ · · · ∪ KN . Then they cover
each individual Kn and a finite number will then suffice in each case,
kn
[
Kn ⊆ Aik .
k=1
For n = 1, . . . , N, the collection of chosen Aik remains finite and together cover all Kn .
□
Remark 1.16
Finiteness of compact is important. For instance, every finite union of closed bounded
intervals is a compact subset of R. But
[
R= [n, n + 1] “countable union of compact sets”.
n∈Z
42
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Theorem 1.20
But f −1 (Ai ) are open sets since f is continuous. Therefore, RHS is an open cover of K. As K
is compact, a finite number of these open sets still cover it,
N
[
f −1 Aik .
K⊆
k=1
Theorem 1.21
Proof. By Theorem 20, since X is compact and f is continuous, f (X) is compact subset of R.
Hence f (X) is closed and bounded subset of R. Since f (X) is bounded it has both a supremum
and an infimum and it is closed, sup f (X) ∈ f (X) and inf f (X) ∈ f (X). Thus there are x0 , x1 ∈ X
such that f (x0 ) = inf f (X) and f (x1 ) = sup f (X), i.e., f attains its minimum and maximum
value at x0 and x1 , respectively. □
43
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Theorem 1.22
Let f : (X, dX ) → (Y, dY ) be a continuous function and (X, dX ) be a compact metric space.
Then f is uniformly continuous.
Proof. Since f : X → Y is continuous, given ε > 0, for every x ∈ X there exists δ x such that
ε
dY ( f (x), f (y)) < for all y ∈ B(x, δ x ). (∗)
2
The collection of balls B(x, δ x /2) x∈X is an open cover of X. Since X is compact, there are
points x1 , . . . , xn such that X = ∪nk=1 B(xk , δ xk /2). Choose 0 < δ < 21 min{δ x1 , . . . , δ xn } and let x, y
be points in X such that dX (x, y) < δ. Then x ∈ B(xi , δ xi /2) for some 1 ≤ i ≤ n and
δ xi δ xi δ xi
dX (y, xi ) ≤ dX (y, x) + dX (x, xi ) < δ + ≤ + = δ xi .
2 2 2
By (∗),
ε ε
dY ( f (x), f (y)) ≤ dY ( f (x), f (xi )) + dY ( f (xi ), f (y)) < + = ε,
2 2
which proves that f is uniformly continuous. □
Note that the statement of Heine-Borel Theorem is not true if we replace Rn by an arbitrary
metric space. For example, take X = (0, 1) with the usual metric d(x, y) = |x − y|. Let
K = X. The set K is closed in X. Also, K is bounded since d(x, y) < 1 for all x, y ∈ K.
However K is not compact.
1. K is compact
2. Every collection of closed subsets of X with finite intersection property has a non-
empty intersection.
44
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
K).
4. K is complete and totally bounded (it can be covered by a finite number of ϵ-balls).
✓ For the standard Euclidean space (Rn , ∥.∥2 ) the Heine-Borel Theorem asserts that a subset
of Rn is compact if and only if it is closed and bounded.
✓ This continues to hold for every finite dimensional normed vector space and conversely,
every normed vector space in which the closed unit ball is compact is finite-dimensional.
✓ For the Banach space of continuous function on a compact metric space a characterization
of the compact subsets is given by Arzéla-Ascoli Theorem.
Remark 1.18
C(X, Y) := { f : X → Y | f is continuous}
X→R
x 7→ dY ( f (x), g(x))
45
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
So given any ε > 0, ∃N such that dY ( fn (x), fm (x)) < 2ε for any n, m ≥ N and any x ∈ X. For
each x, we can choose m ≥ N, depend on x and large enough so that dY ( fm (x), f (x)) < 2ε and
this implies for any x ∈ X,
dY ( fn (x), f (x)) ≤ dY ( fn (x), fm (x)) + dY ( fm (x), f (x)) < ε
for any n ≥ N. Since this N is independent of x, it follows that d( fn , f ) → 0.
Remark 1.19
Definition 1.20
A subset F ⊂ C(X, Y) is called equicontinuous if for every ε > 0, there exists a constant
δ > 0 such that for all x, x′ ∈ X and all f ∈ F ,
Let (X, d) be a compact metric space and let F ⊂ C(X, Rn ). Then F is compact if and only
if it is closed, bounded and equicontinuous.
46
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
A normed space X is a vector space over R or C with a function ∥.∥ : X → R+ (called the
norm) such that for any x, y ∈ X, λ ∈ R or C,
1. ∥x∥ = 0 ⇔ x = 0.
2. ∥λx∥ = |λ|∥x∥.
3. ∥x + y∥ ≤ ∥x∥ + ∥y∥.
Consequences:
1. ∥x − y∥ ≥ ∥x∥ − ∥y∥ .
2. ∥x1 + · · · + xn ∥ ≤ ∥x1 ∥ + · · · + ∥xn ∥.
X is a Banach space if the metric space (X, d) is complete; i.e., if every Cauchy sequence
in X converges, where
d(x, y) := ∥x − y∥.
Remark 1.20
d(x, y) = ∥x − y∥ for x, y ∈ X
defines a distance on X.
Example 1.26
X p 1/p
n
∥x∥ p = |xk | , x = (x1 , . . . , xn ) ∈ Rn .
k=1
47
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Thanks to the Minkowski’s inequality, the space (Rn , ∥.∥ p ) is normed space. It remains to prove
the completeness. We will prove more general result in the next example for the completeness.
Example 1.27
It suffices to prove completeness for the Euclidean norm. The general result follows from
Theorem 28.
Suppose e1 , e2 , . . . , en is a basis of X and let (xk ) be a Cauchy sequence in X. Then for each k,
we can write xk in terms of basis,
Xn
xk = cki ei .
i=1
It follows that
n
X
|cki − cli |2 ≤ |cki − cli |2 = ∥xk − xl ∥22 ,
i=1
This implies that (cki ) is a Cauchy sequence for each i. Let ci be its limit and let x =
Pn
i=1 ci ei .
Then we have
n
X
∥xk − x∥2 =
2
|cki − ci |2 −→ 0,
i=1
Example 1.28
∥ f ∥∞ := sup | f (x)|.
x∈[0,1]
is a Banach space.
Note that X is a vector space and that ∥ f ∥∞ = 0 implies f = 0. The other conditions of norm
are satisfied
48
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Since this inequality holds for all x ∈ [0, 1], it follows that ∥ f − fn ∥∞ < ϵ for all n ≥ N. Hence
fn converges uniformly to f and so f is continuous and f ∈ X.
Example 1.29
It is easy to check that ∥.∥1 is a norm on X. Now, consider the functions given by
0, 0 ≤ x ≤ 12 − 1n ,
fn (x) = nx − 2n + 1, 12 − n1 ≤ x ≤ 12 ,
1
1,
2
≤ x ≤ 1,
for n ≥ 3.
We claim that ( fn ) is a Cauchy seuqence in X. To see this, let m ≥ n ≥ 3, we have
Z 21
1
∥ fn − fm ∥1 = | fn (x) − fm (x)|dx ≤ → 0,
1 1
2−n
n
as n → ∞, thus ( fn ) is a Cauchy sequence with respect to ∥.∥1 .
We claim that ( fn ) does not converge to some f ∈ X. Take any a ∈ 0, 12 and let n ∈ N with
1
2
− 1n ≥ a. We then have
Z a Z a
0≤ | f (x)|dx = | f (x) − fn (x)|dx ≤ ∥ fn − f ∥1 → 0,
0 0
Ra
as n → ∞. This implies that 0 | f (x)|dx = 0. The continuity of f implies that f = 0 on [0, a]
for every a < 12 and hence f 12 = 0. On the other hand,
Z 1 Z 1
0≤ | f (x) − 1|dx = | f (x) − fn (x)|dx ≤ ∥ f − fn ∥1 → 0,
1 1
2 2
49
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Example 1.30
X p 1/p
∞
∥x∥ p := |xk | , x ∈ lp.
k=1
The space (l p , ∥.∥ p ) is a normed space follows from the Minkowski’s inequality. It remains to
see why it is complete.
The case p = ∞. Let (xk ) be a Cauchy sequence in l∞ . Write xk = (xk1 , xk2 , . . .). We have to find
an element x ∈ l∞ such that (xk ) converges to x. Let ϵ > 0. Because (xk ) is a Cauchy sequence,
there exists an integer N > 0 such that
∥xk − xl ∥∞ < ϵ,
for all m and all k, l ≥ N. This implies that for each m, the sequence
x1m , x2m , . . .
lim xkm = xm .
k→∞
Let x denote the sequence x = (x1 , x2 , . . .). We want to prove that xk converges to x. Choose an
ϵ > 0. By replacing ϵ with ϵ/2, there exists an integer N > 0 such that
50
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
It follows that ϵ p
∥xk − x∥ pp <
2
p
for all k ≥ N and xk converges to x and that x ∈ l .
Example 1.31
where f : M → R measurable.
For p = ∞, (L∞ (µ), ∥.∥∞ ) is a Banach space of bounded measurable functions with
Note that
51
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
2. The case X = N and µ equals to the counting measure. Then L p and l p are coincide.
The space (L p (µ), ∥.∥ p ) is a normed space follows from the Minkowski’s inequality. To prove
the completeness we distinguish two cases:
Case 1 ≤ p < ∞:
Step 1. The Cauchy sequence: Let ( fn ) be a Cauchy sequence in L p . It suffices to show that
( fn ) has a convergence subsequence with limit f in L p . By definition of Cauchy sequence, for
each k ∈ N, we can find Nk ∈ N so that ∥ fm − fn ∥ p < 21k whenever m, n ≥ Nk . In addition,
construct an increasing sequence of natural numbers by choosing n1 = N1 and for k ≥ 2, set
nk = max{nk−1 , Nk }+1. Then we have a subsequence ( fnk ) with ∥ fnk+1 − fnk ∥ < 21k for every k ∈ N.
We want to show that ( fnk ) converges to a function f ∈ L p . We simply refer to fnk as fk .
Step 2. Conversion to a convergent monotone sequence. Define a sequence of functions (gk )
by g1 (x) = | f1 (x)| and
It follows that
0 ≤ g1 (x) ≤ g2 (x) ≤ g3 (x) ≤ · · ·
and
∥gk ∥ p = ∥| f1 | + | f2 − f1 | + | f3 − f2 | + · · · + | fk − fk−1 |∥ p
≤ ∥ f1 ∥ p + ∥ f2 − f1 ∥ p + ∥ f3 − f2 ∥ p + · · · + ∥ fk − fk−1 ∥ p
X k
= ∥ f ∥p + ∥ f j − f j−1 ∥ p
j=2
< ∥ f1 ∥ p + 1,
where we used
k ∞ ∞
X X X 1
∥ f j − f j−1 ∥ p < ∥ f j − f j−1 ∥ p < = 1.
j=2 j=1 j=1
2j
This shows that gk ∈ L p for every k ∈ N. Set g(x) = limk→∞ gk (x) and g p (x) = limk→∞ gk (x) p .
Moreover, the sequence (gk (x)) p is increasing for each x ∈ X and an application of monotone
convergence theorem yields
Z !1/p Z !1/p Z !1/p
∥g∥ p = p
|g| dµ = p
lim gk dµ = lim p
gk dµ
X X k→∞ k→∞ X
Z !1/p
= lim p
gk dµ = lim ∥gk ∥ p ≤ ∥ f1 ∥ p + 1.
k→∞ X k→∞
52
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
(gk (x)) is a Cauchy in R for every x. That is, for all ϵ > 0, there exists N ∈ N such that
|gk (x) − gm (x)| = gk (x) − gm (x) < ϵ whenever k > m ≥ N and thus
k
X k
X
| fk (x) − fm (x)| = f j (x) − f j−1 (x) ≤ | f j (x) − f j−1 (x)| = gk (x) − gm (x) < ϵ,
j=m+1 j=m+1
Case p = ∞: Let ( fn ) be a Cauchy sequence in L∞ (µ) and let Ak and Bm,n be the sets such that
| fk (x)| > ∥ fk ∥∞ and | fn (x) − fm (x)| > ∥ fn − fm ∥∞ , repsectively and let E be the union of these
sets for k, m, n = 1, 2, 3, . . . . Then µ(E) = 0 and on the complement of E the sequence ( fn )
converges uniformly to a bounded function f . Define f (x) = 0 for x ∈ E. Then f ∈ L∞ (µ) and
∥ fn − f ∥∞ → 0 as n → ∞.
Remark 1.21
The main ingredients in the proof consist of Minkowski’s inequality, Monotone conver-
gence theorem, Lebesgue dominated convergence theorem and the following fact.
Lemma 1.2
Let (X, d) be a metric space. Then X is complete if and only if every Cauchy sequence
has a convergent subsequence (with limit x ∈ X).
Proof. Let (xn ) be a Cauchy sequence in the metric space (X, d) and (xnk ) be its conver-
gence subsequence, say xnk → x. We claim that (xn ) is convergent with the limit x. Then
∀ϵ > 0, ∃N ∈ N such that for all n, m, k > N,
ϵ ϵ
d(xn , xm ) < and d(xnk , x) < .
2 2
Therefore, note that nk ≥ k, we have
ϵ ϵ
d(xn , x) ≤ d(xn , xnk ) + d(xnk , x) < + = ϵ.
2 2
53
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Let (X, ∥.∥X ) and (Y, ∥.∥Y ) be real normed spaces. A linear operator T : X → Y is bounded
if there exists a constant c ≥ 0 such that
Example 1.32
T ( f ) = f (0)
is bounded.
Indeed, we have
|T ( f )| = | f (0)| ≤ sup {| f (x)|} = ∥ f ∥.
x∈[0,1]
Remark 1.22
Then T is unbounded.
54
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Example 1.33
Note that k is continuous on the compact set [0, 1] × [0, 1] hence it is bounded, that is, ∥k∥∞ =
max x,y∈[0,1] |k(x, y)| < ∞. Thus,
Z 1
∥T f ∥∞ = sup |(T f )(x)| = sup k(x, y) f (y)dy
x∈[0,1] x∈[0,1] 0
Z 1
≤ sup |k(x, y)|| f (y)|dy ≤ ∥ f ∥∞ ∥k∥∞
x∈[0,1] 0
Example 1.34
Let (X, ∥.∥∞ ) be the normed space of all polynomials on [0, 1] with ∥ f ∥∞ = max x∈[0,1] | f (x)|.
A differential operator T is defined on X by (T f )(x) = f ′ (x). Then T is unbounded.
Remark that T is linear operator due to the linearity of differentiation. To see that T is un-
bounded, we consider fn (x) = xn then (T fn )(x) = nxn−1 , x ∈ [0, 1] and
Thus
∥T fn ∥∞ = n and ∥T fn ∥∞ /∥ fn ∥∞ = n → ∞ as n → ∞.
So T is unbounded.
55
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Example 1.35
is bounded.
T is bounded:
∞
X ∞
X ∞
X
|T (xk )| = ak xk ≤ |ak xk | ≤ sup{|ak |} |xk | = ∥a∥∞ ∥x∥1 .
k=1 k=1 k≥1 k=1
Example 1.36
is bounded.
56
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Example 1.37
Hence T m is bounded with c = ∥m∥∞ . It follows that ∥T m ∥L(l p ) ≤ ∥m∥∞ . To see that the equality
holds, consider (ek ) ∈ l p , where ek = (0, . . . , 0, 1, 0, . . .) with a 1 at the k-th position. Then
∥ek ∥ p = 1 and ∥T m ek ∥ p = |mk |. Thus ∥T m ∥L(l p ) ≥ |mk | for all k ∈ N. Hence ∥T m ∥L(l p ) ≥ ∥m∥∞ . The
case p = ∞ is similar.
from the definition of supremum since we are taking suprema of the same quantity over a larger
set. Since
∥T x∥Y x
= T
∥x∥X ∥x∥X
we have
∥T x∥Y
sup ≤ sup ∥T x∥Y .
x,0 ∥x∥X ∥x∥X =1
57
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
If ∥x∥X ≤ 1 we have
∥T x∥Y
∥T x∥Y ≤ .
∥x∥X
Then
∥T x∥Y
sup ∥T x∥Y ≤ sup
∥x∥X ≤1 x,0 ∥x∥X
and thus
∥T x∥Y
sup = sup ∥T x∥Y = sup ∥T x∥Y .
x,0 ∥x∥X ∥x∥X ≤1 ∥x∥X =1
Finally, we have
∥T x∥Y
∥T x∥Y ≤ sup ∥x∥X
x,0 ∥x∥X
for all x ∈ X. Then ∥T ∥ ≤ sup x,0 ∥T∥x∥x∥XY . Moreover, by the definition of supremum there exists a
sequence (xn ) such that
∥T xn ∥Y ∥T x∥Y 1
≥ sup − for all n.
∥xn ∥X x,0 ∥x∥X n
By the definition of ∥T ∥ we have that for all n, ∥T ∥ ≥ ∥T∥xxnn∥∥XY . Therefore, ∥T ∥ = sup x,0 ∥T x∥Y
∥x∥X
.
The space of bounded linear operators from X to Y is denoted by
L(X, Y) := T : X → Y | T is linear and bounded .
Theorem 1.26
Let (X, ∥.∥X ) and (Y, ∥.∥Y ) be real normed spaces. Then (L(X, Y), ∥.∥L(X,Y) ) is a normed
space.
Finally,
∥T + S ∥L(X,Y) = sup ∥(T + S )x∥Y
∥x∥X ≤1
= ∥T ∥L(X,Y) + ∥S ∥L(X,Y) .
It follows that (L(X, Y), ∥.∥L(X,Y) ) is a normed space. □
58
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Theorem 1.27
Let (X, ∥.∥X ) and (Y, ∥.∥Y ) be real normed spaces and let T : X → Y be a linear operator.
TFAE
(1). T is bounded
(2). T is continuous
(3). T is continuous at x = 0.
∥T x − T x′ ∥Y = ∥T (x − x′ )∥Y ≤ ∥T ∥∥x − x′ ∥X
for all x, x′ ∈ X and so T is Lipschitz continuous. Since every Lipschitz continuous function is
continuous, this shows that (1) ⇒ (2).
(2) ⇒ (3): By definition of continuity.
(3) ⇒ (1): Assume that T is continuous at x = 0. Then it follows from ϵ − δ definition of
continuity with ϵ = 1 that there exists a constant δ > 0 such that for all x ∈ X,
∥x∥X ≤ δ ⇒ ∥T x∥Y ≤ 1.
This implies ∥T x∥Y ≤ 1 for every x ∈ X with ∥x∥X ≤ δ. Now let x ∈ X \ {0}. Then
X x∥X ≤ δ
∥δ∥x∥−1
and so
T δ∥x∥−1
X x ≤1
Y
Hence
∥T x∥Y ≤ δ−1 ∥x∥X , for all x ∈ X.
□
∥T x∥ ≤ ∥T ∥∥x∥.
59
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Definition 1.24
Let X be a real vector space. Two norms ∥.∥ and ∥.∥′ on X are equivalent if there is a
constant C ≥ 1 such that
1
∥x∥ ≤ ∥x∥′ ≤ C∥x∥ for all x ∈ X.
C
Remark 1.23
Definition 24 is equivalent to
1 ′
∥x∥ ≤ ∥x∥ ≤ C∥x∥′ for all x ∈ X.
C
An alternative condition for the norms equivalent is that there exist positive real numbers
60
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
c, d such that
c∥x∥ ≤ ∥x∥′ ≤ d∥x∥ for all x ∈ X.
and this condition is equivalent to the above by setting C = max{1/c, d}.
Theorem 1.28
Let X be a finite dimensional real vector space. Then any two norms on X are equivalent.
This is a norm on X. We prove that every norm on X is equivalent to ∥.∥2 . Fix any norm function
X → R : x 7→ ∥x∥.
Step 1: There is a constant c > 0 such that ∥x∥ ≤ c∥x∥2 for all x ∈ X. Define
n 12 n
X X
x= xi ei , xi ∈ R.
c := ∥ei ∥2 and let
i=1 i=1
Then by triangle inequality for ∥.∥ and the Cauchy-Schwarz inequality on Rn , we have
n
n 1 n 12
X X 2 2 X
∥ei ∥ ≤ c∥x∥2 .
2
∥x∥ ≤ |xi |∥ei ∥ ≤ |xi |
i=1 i=1 i=1
Step 2: There is a constant δ > 0 such that δ∥x∥2 ≤ ∥x∥ for all x ∈ X.
The set S := x ∈ X | ∥x∥2 = 1 is compact with respect to ∥.∥2 by the Heine-Borel Theorem
23 and the function S → R : x 7→ ∥x∥ is continuous. Hence by Theorem 21, there exists an
element x0 ∈ S such that ∥x0 ∥ ≤ ∥x∥ for all x ∈ S . Define δ := ∥x0 ∥ > 0. The every nonzero
2 x ∈ S , hence ∥x∥2 x ≥ δ and so ∥x∥ ≥ δ∥x∥2 .
vector x ∈ X satisfies ∥x∥−1 −1
□
Corollary 1.3
Proof. This holds for Rn for Euclidean norm. By Theorem 28 it holds for every norms on Rn .
Thus it holds for every finite dimensional normed spaces. □
61
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Corollary 1.4
Proof. Let Y ⊂ X be a finite dimensional linear subspace and denote by ∥.∥Y the restriction of
the norm on X to the subspace Y. Then (Y, ∥.∥Y ) is complete and hence Y is closed subset of
X. □
Corollary 1.5
Let (X, ∥.∥) be a finite dimensional normed space and let K ⊂ X. Then K is compact if and
only if K is closed and bounded.
Proof. This holds for Euclidean norm on Rn by Heine-Borel Theorem 23. Hence it holds for
every norms on Rn . Hence it holds for every finite dimensional normed spaces. □
Theorem 1.29
Let (X, ∥.∥X ) be a finite dimensional normed space and let (Y, ∥.∥Y ) be any normed space.
Then any linear operator T : X → Y is bounded.
Theorem 1.30
Let (X, ∥.∥) be a normed space and denote the closed unit ball and the closed unit sphere in
X by
B := {x ∈ X | ∥x∥ ≤ 1}, S := {x ∈ X | ∥x∥ = 1}.
Then the following are equivalent
(2). B is compact
62
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
(3). S is compact.
Let (X, ∥.∥) be a normed space and let Y ⊂ X be a closed linear subspace that is not equal
to X. Fix a constant 0 < δ < 1. Then there exists a vector x ∈ X such that
∥x∥ = 1, inf ∥x − y∥ ≥ 1 − δ.
y∈Y
Proof. Let x0 ∈ X \ Y. Then d := inf y∈Y ∥x0 − y∥ > 0 because Y is closed. Choose y0 ∈ Y such
d
that ∥x0 − y0 ∥ ≤ 1−δ and define x = ∥x0 − y0 ∥−1 (x0 − y0 ). Then ∥x∥ = 1 and
x0 − y0 − ∥x0 − y0 ∥y d
∥x − y∥ = ≥ ≥1−δ
∥x0 − y0 ∥ ∥x0 − y0 ∥
for all y ∈ Y. □
63
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Theorem 1.31
Let X be a normed space and let Y be a Banach space. Then L(X, Y), ∥.∥L(X,Y) is a Banach
space with respect to the operator norm.
for all x ∈ X and all m, n ∈ N. Hence (T n x)n∈N is a Cauchy sequence in Y for every x ∈ X. Since
Y is complete, this implies that the limit T x := limn→∞ T n x exists for all x ∈ X. This defines a
map T : X → Y.
• T is linear since
T (x + y) = lim T n (x + y) = lim (T n x + T n y)
n→∞ n→∞
= lim T n x + lim T n y = T x + T y
n→∞ n→∞
and
• T is bounded since
∥T x∥ = lim ∥T n x∥ ≤ C∥x∥.
n→∞
• We now prove that limn→∞ T n = T . Let ϵ > 0. Since (T n )n∈N is a Cauchy sequence with
repsect to the operator norm, there exists an N ∈ N such that when m, n ≥ N,
ϵ
∥T n − T m ∥ < .
2
Then for any x with ∥x∥X ≤ 1 and any m, n ≥ N,
ϵ
∥T n x − T m x∥Y ≤ ∥T n − T m ∥∥x∥X < .
2
Since T x = limn→∞ T n x, there exists an N1 ∈ N such that when m ≥ N1 ,
ϵ
∥T x − T m x∥Y < .
2
64
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Remark 1.24
In particular,
• if X is a Banach space then the space L(X) := L(X, X) of bounded linear operators
from X to itself is a Banach space.
• for any normed space (X, ∥.∥X ) the dual space X ∗ = L(X, R) (or X ∗ = L(X, C) if X is
a complex vector space) is complete as both R and C are complete. (For more details
see Section 1.6).
H × H → C : (x, y) 7→ ⟨x, y⟩
is called an inner product if for any x, y, z ∈ H and α, β ∈ C the following conditions are
satisfied
65
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Remark 1.25
By Definition 25, the inner product ⟨x, y⟩ of two vectors is a complex number. By (i).
⟨x, x⟩ = ⟨x, x⟩, it follows that ⟨x, x⟩ is a real number for every x ∈ H. The condition (ii).
implies that
In particular,
⟨αx, y⟩ = α⟨x, y⟩ and ⟨x, αy⟩ = ᾱ⟨x, y⟩.
Every inner product space is also a normed space with the norm defined by
p
∥x∥ = ⟨x, x⟩.
Indeed, this functional is well defined because ⟨x, x⟩ is always a nonnegative real number and
∥x∥ = 0 if and only if x = 0. In addition,
p q
∥λx∥ = ⟨λx, λx⟩ = λλ̄⟨x, x⟩ = |λ|∥x∥.
It is left with the tirangle inequality where we use the Cauchy-Schwarz inequality.
Lemma 1.4
66
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Propoistion 1.19
Proof. Let xn → x and yn → y. Since a convergence sequence implies Cauchy sequence and
Cauchy sequence implies bounded sequence hence yn is bounded. We have
|⟨xn , yn ⟩ − ⟨x, y⟩| ≤ |⟨xn − x, yn ⟩| + |⟨x, yn − y⟩|
≤ ∥xn − x∥∥yn ∥ + ∥x∥∥yn − y∥ → 0
□
67
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Let (X, M, µ) be a measure space then L2 (µ) is a Hilbert space with the inner product
Z
⟨ f, g⟩ = f ḡdµ.
X
Indeed, the norm associated to the given inner product is the L2 -norm.
sZ
∥ f ∥ = ⟨ f, f ⟩ = | f |2 dµ = ∥ f ∥2
p
X
We have already proved that L2 (µ) with the L2 -norm is complete and hence L2 (µ) is a
Hilbert space.
In the case where X = N, it follows that l2 is a Hilbert space with the inner product
∞
X
⟨x, y⟩ = xn ȳn ,
n=1
Do all norms on vector spaces come from inner products and if not, which property character-
izes inner product spaces?
A norm is induced from an inner product if and only if it satisfies, for all x, y,
The statement asserts that the sum of the lengths squared of diagonals of parallelogram equals
that of the sides.
68
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Proof. (⇒) In any Hilbert space the parallelogram identity holds. It follows from adding the
identities
∥x + y∥2 = ∥x∥2 + 2Re⟨x, y⟩ + ∥y∥2
∥x − y∥2 = ∥x∥2 − 2Re⟨x, y⟩ + ∥y∥2 .
(⇐) If the parallelogram indentity holds in a Banach space then its norm comes from an inner
product which can be recovered in terms of the norm by the polarization identity: for complex
cases
1
⟨x, y⟩ = ∥x + y∥2 − ∥x − y∥2 + i∥x + iy∥2 − i∥x − iy∥2
4
and for real spaces
1
⟨x, y⟩ = ∥x + y∥2 − ∥x − y∥2 .
4
□
Theorem 1.32
Let H be a Hilbert space and let K ⊂ H be a nonempty closed convex subset of H. Then
there exists a unique element x0 ∈ K such that ∥x0 ∥ ≤ ∥x∥ for all x ∈ K.
69
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Let H be a Hilbert space and let Λ : H → R be a bounded linear functional. Then there
exists a unique element y ∈ H such that
Proof. Existence:
If Λ = 0 then the vector y = 0 satisfies (1.5.1). Hence assume Λ , 0 and define
K := {x ∈ H | Λ(x) = 1}.
Then K , ∅ because there exists an element ξ ∈ H such that Λ(ξ) , 0 and hence x := Λ−1 (ξ)ξ ∈
K. The set K is closed because Λ : H → R is continuous and it is convex because Λ is linear.
Hence Theorem 32 asserts that there exists an element x0 ∈ K such that ∥x0 ∥ ≤ ∥x∥ for all
x ∈ K. We claim that
x ∈ H, Λ(x) = 0 ⇒ ⟨x0 , x⟩ = 0.
Indeed, let x ∈ H such that Λ(x) = 0. Then x0 + tx ∈ K for all t ∈ R. This implies for all t ∈ R,
Thus the differentiable function t 7→ ∥x0 + tx∥2 attains its minimum at t = 0 and so its derivative
vanishes at t = 0. Hence
d
0= ∥x0 + tx∥2 = 2⟨x0 , x⟩.
dt t=0
This proves that ⟨x0 , x⟩ = 0.
Now define y = ∥x0 ∥−2 x0 . Let x ∈ H and define λ = Λ(x). Then
70
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Corollary 1.6
71
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Definition 1.27
Let X be a normed space over R (or C). The space L(X, R) (or L(X, C)) is called the dual
space of X and is denoted by X ∗ .
Remark 1.26
Dual space of Rn
Theorem 1.34
∗
There is an isometric isomorphism map T : Rn → Rn that sends each vector α to the
bounded linear operator T α defined by
n
X
T α (x) = αi xi .
i=1
Proof. The operator T α is linear for each α ∈ Rn . To see that T α is bounded, we apply the
Cauchy-Schwarz inequality
n
X n
X n
1/2 X 1/2
|T α (x)| ≤ |αi ||xi | ≤ α2i xi2 = ∥α∥2 ∥x∥2 .
i=1 i=1 i=1
This implies that ∥T α ∥ ≤ ∥α∥2 for each α ∈ Rn . It follows that T α is both bounded and linear,
∗
so it is an element of the dual Rn . Since the equality is attained for x = (α1 , . . . , αn ), we
∗
have ∥T α ∥ = ∥α∥2 . To see that T is an isomorphism, consider a map T : Rn → Rn defined
by T (α) = T α . First we show that it is injective. Suppose that T α = T β for α, β ∈ Rn . Then
T α (ei ) = T β (ei ) for each i where (e1 , . . . , en ) is the standard basis of Rn . Thus αi = βi for each i.
72
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
It follows that α = β. Then we prove that it is surjective. Suppose S is an element of the dual
∗
Rn and consider the vector
α = S (e1 ), . . . , S (en ) ∈ Rn .
Theorem 1.35
Then Λy ∈ H ∗ and
∥Λy ∥ = ∥y∥.
The map
Λ : H → H∗
y 7→ Λy
is an isometric isomorphism.
73
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Let (M, F , µ) be a measure space and let 1 < p < ∞. Define the number 1 < q < ∞ by
1 1
+ = 1.
p q
Let g ∈ Lq (µ) and define Λg : L p (µ) → R by
Z
Λg ( f ) = f gdµ for f ∈ L p (µ).
M
Λ : Lq (µ) → L p (µ)∗
g 7→ Λg
is an isometric isomorphism.
The key idea of the proof: The Hölder’s inequality asserts that f ∈ L p (µ), g ∈ Lq (µ) then
f g ∈ L1 (µ) and Z
|Λg ( f )| = f gdµ ≤ ∥ f ∥ p ∥g∥q .
M
It turns out that Λ ∈ L p (µ)∗ and
∥Λg ∥ ≤ ∥g∥q
for all g ∈ Lq (µ). On the other hand, the function f1 = |g|q−2 g satisfies
Z Z Z
| f1 | dµ =
p
|g|(q−1)p
dµ = |g|q dµ < ∞
M M M
74
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Remark 1.27
If (M, F , µ) be σ-finite. The dual of L1 (µ) is L∞ (µ). This may not hold when the measure
is not σ-finite. On the other hand, the dual space of L∞ (µ) is general larger than L1 (µ).
Dual space of l p
Theorem 1.37
∥T y ∥ = ∥y∥q .
Now ∞ ∞
X X
T y (x) = xk yk = |yk |q = ∥y∥qq .
k=1 k=1
Thus
|T y (x)| ∥y∥qq
∥T y ∥ ≥ = q/p
= ∥y∥q−q/p
q = ∥y∥q .
∥x∥q ∥y∥q
For more details see [2].
75
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
76
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Chapter 2
Main Theorems
• Uniform boundedness principle: every pointwise bounded family of bounded linear op-
erators on a Banach space is bounded.
• Open mapping theorem: every surjective bounded linear operator between two Banach
spaces is open.
• Closed graph theorem: a linear operator between two Banach spaces is bounded if and
only if its graph is a closed linear subspace of the product space.
77
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Proof. Assume that Åk = ∅ for all k. Then for U ⊂ Ak open, nonempty, U \ Ak is open,
nonempty for any k ∈ N. Then there exists a ball Bε (x) ⊂ U \ Ak with ε ≤ 1k . Hence we can
inductively choose balls Bεk (xk ) such that
1
Bεk (xk ) ⊂ Bεk−1 (xk−1 ) \ Ak and εk ≤ .
k
Consequently, we see that xl ∈ Bεk (xk ) for l ≥ k and εk → 0 as k → ∞ and the balls Bεk (xk ) are
nested and we conclude that (xl )l∈N is a Cauchy sequence. Hence there exists the limit
x = lim xl ∈ X
l→∞
and x ∈ Bεk (xk ) for all k. Since Bεk (xk ) ∩ Ak = ∅, we have that x < Ak = X, a contradiction.
S
k∈N
□
This is just one of a couple forms of Baire’s Category theorem. Another version is that the
intersection of countably many open dense sets is dense.
Definition 2.1
Theorem 2.2
Let (X, d) be a complete metric space. Then the intersection of every countable collection
of dense open subsets of X is dense in X.
Proof. The basic idea of the proof: let V1 , V2 , . . . be dense and open in X and let W be any open
T
set in X. We have to show that ( Vk ) ∩ W , ∅.
Since V1 is dense, W ∩ V1 is a nonempty set, and we can therefore find x1 and r1 such that
If k ≥ 2 and xk−1 and rk−1 are chosen, the denseness of Vk shows that Vk ∩ B(xk−1 , rk−1 ) is not
empty and we can therefore find xk and rk such that
1
B̄(xk , rk ) ⊂ Vk ∩ B(xk−1 , rk−1 ) and 0 < rk < .
k
By induction, this process produces a sequence (xk ) in X. If i > k and j > k, the construction
shows that xi and x j both lie in B(xk , rk ) so that d(xi , x j ) < 2rk < 2k , and hence (xk ) is a Cauchy
78
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Theorem 2 is equivalent to the statement that no complete metric space is of the first category.
To see this, just take complements in the statement of Theorem 2.
Let X be a nonempty complete metric space and let Y be a normed space. Let F ⊂ C(X, Y)
be a set of functions with
Proof. For f ∈ F and k ∈ N, it holds that x ∈ X | ∥ f (x)∥Y ≤ k is a closed set (as f is
continuous). Hence the sets
\
Ak = x ∈ X | ∥ f (x)∥Y ≤ k
f ∈F
being intersection of closed sets, are closed and it follows from (∗) that they form a cover of
X. Then Baire’s Category Theorem 1 yields that Åk , ∅ for some k0 and hence there exists a
Bε0 (x0 ) ⊂ Ak0 . Noting that
sup sup ∥ f (x)∥Y < k0
x∈Ak0 f ∈F
Now we present the Banach-Steinhaus theorem which states that every pointwise bounded
family of bounded linear operators on a Banach space is bounded.
79
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Let X be a Banach space and let Y be a normed space. Suppose T ⊂ L(X, Y) with
We remark that there is a notion of convergence in the space L(X, Y) of bounded linear
operators from X to Y in operator norm. A sequence (T n ) in L(X, Y) is uniformly convergent to
T ∈ L(X, Y) if ∥T n − T ∥ → 0 as n → ∞. (In other words, uniform convergence of a sequence
of operators is simply convergence in the operator norm).
A weaker notion of convergence for a sequence of operators is given by the following
definition.
Definition 2.2
Corollary 2.1
80
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
strongly convergent, then there exists an operator T ∈ L(X, Y) such that (T n ) is strongly
convergent to T .
Theorem 2.5
This theorem states that the nth partial sum of Fourier series
n
X
(S n f )(x) = fˆ(k)e2πikx
k=−n
Indeed,
n Z
X
(S n f )(x) = f (s)e−2πiks ds e2πikx
k=−n T
n
Z X
= e2πi(x−s)
f (s)ds
T k=−n
Z
= Dn (x − s) f (s)ds.
T
81
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
The Dirichlet kernel Dn is real valued and can be expressed in the form
1. n
X
Dn (x) = 1 + 2 cos(2πkx).
k=1
2.
2n + 1, if x = 0
Dn (x) =
n+ 21 2πx
sin sin(πx)
, if x , 0.
and satisfies Z
Dn (x)dx = 1.
T
In particular, we have
Z Z
(S n f )(0) = f (s)Dn (−s)ds = f (s)Dn (s)ds.
T T
Lemma 2.2
is bounded with Z
∥T n ∥ = |Dn (s)|ds.
T
Thus, Z
∥T n ∥ ≤ |Dn (s)|ds.
T
Fix δ > 0. Since Dn is analytic it can only have finitely many sign changes in [0, 1]. Therefore,
we can find a continuous function fn with ∥ fn ∥∞ ≤ 1 that differes from sign(Dn (s)) only a finite
82
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
union of intervals whose total length is less than ∥Dδn ∥∞ . The triangle inequality for intergrals
yields Z Z
fn (s)Dn (s) > |Dn (s)|ds − 2δ,
T T
as δ > 0 was arbitrary, this completes the proof. □
Lemma 2.3
sin n + 21 2πs
Z Z
|Dn (s)|ds = ds → ∞
T T sin(πs)
as n → ∞.
Now | sin x| ≥ 1
2
for all x ∈ πZ + [ π6 , 5π
6
]. In particular, it follows that if
2n h
[ i
(2n + 1)πx ∈ k + 1/6 π, k + 5/6 π
k=0
then
1
| sin((2n + 1)πx)| ≥ .
2
Thus
sin n + 12 2πs 2n Z (k+ 5 )π/(2n+1)
Z X 6 1 1
ds ≥ ds
T sin(πs) 1
k=0 (k+ 6 )π/(2n+1)
π(k + 6 )/(2n + 1) 2
5
1 X 2n + 1 6 π
2n 4
=
2π k=0 k + 56 2n + 1
2n
1X 1
= →∞
3 k=0 k + 5
6
as n → ∞. □
Proof of Theorem 5. By Lemma 2 we have T n f = (S n f )(0) for all f ∈ C(T). Moreover, for
f ∈ C(T), if the Fourier series of f converges at 0, then the family {T n f }n≥1 is bounded. Thus if
the Fourier series of f converges at 0 for all f ∈ C(T), then for each f ∈ C(T) the set {T n f }n≥1
is bounded. By Theorem 4, this implies that the set {∥T n ∥}n≥1 is bounded, which contradicts
Lemma 2 and 3. It follows that there must be some f ∈ C(T) whose Fourier series does not
converge at 0 (and indeed the partial sums must be unbounded). □
83
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
• If X and Y are normed spaces and T : X → Y is linear then T is open if and only if there
exists a δ > 0 with
Bδ (0) ⊂ T (B1 (0))
i.e.,
y ∈ Y | ∥y∥Y < δ ⊂ T x | x ∈ X, ∥x∥X < 1 .
Indeed, let U be open and let x ∈ U. Choose an ε > 0 with Bε (x) ⊂ U. Now, Bδ (0) ⊂
T (B1 (0)) implies Bεδ (T x) ⊂ T (Bε (x)) ⊂ T (U) and hence T (U) is open.
Let X and Y be Banach spaces. Then it holds for every operator T ∈ L(X, Y) that
T is surjective ⇐⇒ T is open.
That is, every surjective bounded linear operator between two Banach spaces is open.
It follows from the Baire’s Category Theorem 1 that there exists a k0 and a ball Bε0 (y0 ) in Y
with
Bε0 (y0 ) ⊂ T (Bk (0)).
This means that for y ∈ Bε0 (0) there exists points xi ∈ Bk0 (0) with T xi → y0 + y as i → ∞. On
choosing an x0 ∈ X with T x0 = y0 , this implies that
!
xi − x0 y xi − x0
T → and < 1,
k0 + ∥x0 ∥ k0 + ∥x0 ∥ k0 + ∥x0 ∥
which proves that
ε0
Bδ (0) ⊂ T (B1 (0)) with δ = . (∗).
k0 + ∥x0 ∥
However, we want to show that such an inclusion holds without the closure of the set on the
right hand side of (∗), for a smaller δ if necessary. We note that (∗) implies that y ∈ Bδ (0)
implies there exists an x ∈ B1 (0) with y − T x ∈ Bδ/2 (0). Then 2(y − T x) ∈ Bδ (0). Hence for
y ∈ Bδ (0) we can inductively choose points yk ∈ Bδ (0) and xk ∈ B1 (0) such that
84
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Then
2−k−1 yk+1 = 2−k yk − T (2−k xk ),
and so m
X −k
T 2 xk = y − 2−m−1 ym+1 → y as m → ∞.
k=0
Since m m
X X
2 xk <
−k
2−k ≤ 2 < ∞,
k=0 k=0
Pm
We have that k=0 2−k xk is a Cauchy sequence in X. Because X is complete, there exists
m∈N
x = k=0 2 xk in X with ∥x∥ < 2. The continuity of T then yields that
P∞ −k
m
X −k
T x = lim T 2 xk = y.
m→∞
k=0
This shows that Bδ (0) ⊂ T (B2 (0)) or equivalently Bδ/2 ⊂ T (B1 (0)). Hence T is open.
(⇐) The fact that Bδ (0) ⊂ T (B1 (0)) for some δ > 0 implies that BR (0) ⊂ T (BR/δ (0)) for all
R > 0. □
That is, every bounded linear operator between two Banach spaces has a bounded inverse.
85
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Definition 2.4
Let X and Y be normed spaces and T : D(T ) → Y a linear operator with domain D(T ) ⊂ X.
Then T is called a closed linear operator if its graph
When X and Y are Banach spaces, X × Y is a Banach space with the norm ∥(x, y)∥. Indeed, let
(zn ) be Cauchy in X × Y, where zn = (xn , yn ). Then for every ε > 0 there is an N such that for
all m, n ≥ N,
∥zn − zm ∥ = ∥xn − xm ∥ + ∥yn − ym ∥ < ε, (∗)
Hence (xn ) and (yn ) are Cauchy in X and Y, respectively and converge, say, xn → x and yn → y
since X and Y are complete. This implies that zn → z = (x, y) since from (∗) with m → ∞ we
have ∥zn − z∥ < ε for n > N. Since the Cauchy sequence (zn ) was arbitrary, X × Y is complete.
Let X and Y be Banach spaces and T : D(T ) → Y a closed linear operator, where D(T ) ⊂
X. Then if D(T ) is closed in X, the operator T is bounded.
Proof. By assumption, G(T ) is closed in X × Y and D(T ) is closed in X. Hence G(T ) and D(T )
are complete. We now consider the mapping
P : G(T ) → D(T )
(x, T x) 7→ x.
• P is bounded since
86
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Since G(T ) and D(T ) are complete, we can apply the inverse mapping theorem and conclude
that P−1 is bounded, say, ∥(x, T x)∥ ≤ b∥x∥ for some b and all x ∈ D(T ). Hence T is bounded
because
∥T x∥ ≤ ∥T x∥ + ∥x∥ = ∥(x, T x)∥ ≤ b∥x∥
for all x ∈ D(T ). □
p(αx) = αp(x)
p(x + y) ≤ p(x) + p(y)
for all x, y ∈ X and for all scalars α ≥ 0. Let Y be a subspace of X and let f : Y → R be a
linear map such that
f (x) ≤ p(x) for all x ∈ Y.
Then there exists a linear extension f˜ : X → R of f (i.e. f˜(x) = f (x) for all x ∈ Y) such
that
f˜(x) ≤ p(x) for all x ∈ X.
87
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
y ∈ Yα ∩ Yβ , (Yα , gα ) ∈ N, (Yβ , gβ ) ∈ N
Then
(Yα , gα ) ≤ (Yβ , gβ ) or (Yβ , gβ ) ≤ (Yα , gα )
and in the first case,
Yα ⊂ Yβ and gα = gβ on Yα
Thus
gα (y) = gβ (y).
It holds true that g = f on Y and g ≤ p on Y ′ . We can show that Y ′ is a subspace and g′ is
′ ′
linear as follows: for all x, y ∈ Y ′ , λ ∈ R. Then there exists (Yα , gα ) ∈ N, (Yβ , gβ ) ∈ N with
x ∈ Yα and y ∈ Yβ . It follows that
This implies x, y ∈ Yξ with ξ = β in the first and ξ = α in the second case, hence also x + λy ∈
Yξ ⊂ Y ′ and
g′ (x + λy) = gξ (x + λy) = gξ (x) + λgξ (y) = g′ (x) + λg′ (y).
Applying Zorn’s Lemma: Indeed, by Zorn’s lemma there is a maximal element (Y0 , g0 ) in M.
We claim that Y0 = X (and we may take f˜ to be g0 ).
Extending by one dimension: Otherwise assume by contradiction that x ∈ X \ Y0 and let Y1 =
span{Y0 , x}. Each element z ∈ Y1 may be expressed uniquely in the form z = y + λx with y ∈ Y0
and λ ∈ R (as x is assumed not to be in the subspace Y0 ). Define a linear function g1 on Y1 by
g1 (y + λx) = g0 (y) + λc, where the constant c will be chosen to guarantee that g1 is bounded by
p. Now if y1 , y2 ∈ Y0 it follows that
Thus
−p(−x − y2 ) − g0 (y2 ) ≤ p(y1 + x) − g0 (y1 ).
This implies
88
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Choose c to be any number in the interval [A, B]. Then, by construction of A and B,
and
−p(−x − y) − g0 (y) ≤ c (2.5.2)
for all y ∈ Y0 . To prove the required bound for g1 we consider different cases as follows:
Therefore, we get
g1 (y + λx) = g0 (y) + λc ≤ p(y + λx)
for all λ ∈ R and y ∈ Y0 .
A contradiction: Thus we have found (Y1 , g1 ) ∈ M with (Y0 , g0 ) ≤ (Y1 , g1 ) and Y0 , Y1 . This
contradicts the maximality of (Y0 , g0 ) and hence f˜ = g0 is defined on all of X and satisfies the
conclusion of the theorem. □
The functional p is the statement of the Hahn-Banach Theorem 9 is called a sublinear func-
tional.
Definition 2.5
and
p(αx) = αp(x)
for all x, y ∈ X and all α ≥ 0, is called a sublinear functional.
Next we present a version of the Hahn-Banach theorem for vector spaces over C.
89
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
p(αx) = |α|p(x)
p(x + y) ≤ p(x) + p(y)
for all x, y ∈ X and for all scalars α. Let Y be a subspace of X and let f : Y → C be a linear
map such that
| f (x)| ≤ p(x) for all x ∈ Y.
Then there exists a linear extension f˜ : X → C of f (i.e. f˜(x) = f (x) for all x ∈ Y) such
that
| f˜(x)| ≤ p(x) for all x ∈ X.
Proof. Let X be a vector space over C. We split a linear map f : Y → C into real and imaginary
parts
f (x) = Re f (x) + iIm f (x)
where Re f (x) and Im f (x) are R-linear. By the linearity of f we have
0 = i f (x) − f (ix) = iRe f (x) − Im f (x) − Re f (ix) − iIm f (ix)
= −(Re f (ix) + Im f (x)) + i(Re f (x) − Im f (ix))
for all x ∈ Y. In particular, Im f (x) = −Re f (ix) for all x ∈ Y. Therefore,
f (x) = Re f (x) − iRe f (ix), (∗)
ofr all x ∈ Y. We now consider X as a vector space over R. We denote that vector space by XR .
Let YR be a subspace of XR . Now Re f (x) is a R-linear functional and by assumption
Re f (x) ≤ | f (x)| ≤ p(x)
for all x ∈ YR . By Theorem 9 there exists a linear extension g̃ : XR → R such that g̃|YR = Re f
and g̃(x) ≤ p(x) for all x ∈ XR . We set
f˜(x) = g̃(x) − ig̃(ix)
for all x ∈ XR so that f˜|Y = f by (∗). To show that f˜ is linear from X → C we only need to look
at multiplication by i as f˜ is linear over R. We have
f˜(ix) = g̃(ix) − ig̃(2 x) = g̃(ix) − ig̃(−x) = g̃(ix) + ig̃(x) = i(g̃(x) − ig̃(ix)) = i f˜(x)
for all x ∈ X. It remains to show that | f˜(x)| ≤ p(x). Let x ∈ X and a real number θ such that
| f˜(x)| = eiθ f˜(x). Then since | f˜(x)| ∈ R and by the definition of f˜, we get
| f˜(x)| = eiθ f˜(x) = f˜ eiθ x = g̃ eiθ x ≤ p eiθ x = eiθ p(x) = p(x).
90
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Before stating the version of the Hahn-Banach theorem for linear functionals, we first indi-
cate how to pass from a complex normed space to a real normed space and vice versa.
Lemma 2.5
f (x) = Re f (x) + iIm f (x) = u(x) − iRei f (x) = u(x) − iRe f (ix) = u(x) − iu(ix).
Conversely, suppose u is R-linear and f (x) = u(x)−iu(ix) for all x ∈ X. Then Re f (x) = u(x)
for all x ∈ X and it follows that f is R-linear. Since
Hence ∥u∥X∗ = ∥ f ∥X∗ . Similarly, if u is bounded, then f is bounded and ∥ f ∥X∗ = ∥u∥X∗ . □
Now we are ready to state and prove the Hahn-Banach theorem for normed spaces as fol-
lows:
91
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Let X be a real or complex normed space and Y a linear subspace. Let f be bounded
linear functional on a subspace Y of a normed space X. There there exists a bounded linear
functional f˜ on X such that f˜(y) = f (y) for all y ∈ Y and
∥ f˜∥X∗ = ∥ f ∥Y ∗
where
∥ f˜∥X∗ = sup | f˜(x)| and ∥ f ∥Y ∗ = sup | f (x)|.
x∈X x∈Y
∥x∥=1 ∥x∥=1
Theorem 11 states that every bounded (continuous) linear functional on a linear space of a
normed space extends to a bounded (continuous) linear functional on the entire normed space.
Proof. The Real Case: Assume first that X is a real normed space. Define p(x) = ∥ f ∥Y ∗ ∥x∥ for
all x ∈ X. We see that p satisfies
and
p(x + y) = ∥ f ∥Y ∗ ∥x + y∥ ≤ ∥ f ∥Y ∗ (∥x∥ + ∥y∥) = p(x) + p(y).
Hence we can apply the Theorem 9 and conclude that there exists a linear functional f˜ on X
such that f˜(y) = f (y) for all y ∈ Y and satisfies
Further, since
− f˜(x) = f˜(−x) ≤ p(−x) = ∥ f ∥Y ∗ ∥ − x∥ = ∥ f ∥Y ∗ ∥x∥ = p(x)
for all x ∈ X, we see that | f˜(x)| ≤ ∥ f ∥Y ∗ ∥x∥ for all x ∈ X. Hence f˜ is bounded on X and
Since f˜(y) = f (y) for all y ∈ Y, it follows that ∥ f˜∥X∗ ≥ ∥ f ∥Y ∗ . Hence, we can conclude that
∥ f˜∥X∗ = ∥ f ∥Y ∗ .
The Complex Case: Now let X be a complex normed space, let Y ⊆ X be a complex linear
subspace. Let f ∈ Y ∗ and define a real linear functional u by u(y) = Re f (y) for y ∈ Y. Let
ũ : X → R be an extension of u with ∥u∥Y ∗ = ∥ũ∥X∗ = ∥ f ∥Y ∗ (by real case above and Lemma 5).
Define f˜ : X → C by f˜(x) = ũ(x) − iũ(ix) for all x ∈ X. By Lemma 5, f is bounded C-linear
functional on X satisfies
92
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Now let see the consequences of the Hahn-Banach theorem. One of the main consequences
is that the dual of a normed linear space is well equipped with functionals as in the following
proposition.
Propoistion 2.1
Let X be a normed space and x ∈ X a non-zero vector. Then there exists f ∈ X ∗ such that
∥ f ∥X∗ = 1 and f (x) = ∥x∥.
Proof. Let Y be the one-dimensional spanned by x. Define g(αx) = α∥x∥ for all scalars α.
Setting p(y) = ∥y∥ we have
for all scalars α. Hence there exists f ∈ X ∗ such that f (αx) = g(αx) for all α and | f (y)| ≤ p(y) =
∥y∥ for all y in X. Hence f ∈ X ∗ and ∥ f ∥X∗ ≤ 1. Since | f (x)| = ∥x∥ be definition of f we have
∥ f ∥X∗ = 1 and f (x) = g(x) = ∥x∥. □
Propoistion 2.2
Let X be a normed space and Y ⊂ X a proper closed subspace. Then for every x0 ∈ X \ Y
there exists f ∈ X ∗ such that f|Y = 0 and f (x0 ) = 1.
g(y + αx0 ) = α,
for all scalars α and y ∈ Y. Then g(x0 ) = 1 and g(y) = 0 for all y ∈ Y. Moreover, since Y is
closed d := dist(x0 , Y) > 0 and
Hence
|g(y + αx0 )| = |α| ≤ d−1 ∥y + αx0 ∥
for all y ∈ Y and all scalars α. Hence g ∈ Y1∗ . Now by Theorem 11 we can conclude that there
exists f ∈ X ∗ such that f|Y = 0 and f (x0 ) = 1. □
93
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Propoistion 2.3
Proof. We observe that | f (x)| ≤ ∥ f ∥X∗ ∥x∥ ≤ ∥x∥ when ∥ f ∥X∗ ≤ 1. On the other hand, by
Proposition 1, there exists f ∈ X ∗ such that ∥ f ∥X∗ = 1 and f (x) = ∥x∥ when x is non-zero. So
the result holds for non-zero vectors and is trivially true for zero vector. □
∥J x ∥X∗∗ = ∥x∥X .
Definition 2.6
Remark 2.1
1. Because X ∗∗ , being the dual space, is always complete, the notion of reflexivity
makes sense only for Banach spaces.
2. Since dual spaces are always complete, every reflexive space is Banach space.
3. Not every Banach space is reflexive. Example of non-reflexive spaces are l1 , l∞ , L1 (µ)
and L∞ (µ).
94
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Example 2.1
Example 2.2
Let X be a Banach space. Then X is reflexive if and only if the unit ball
B = {x ∈ X | ∥x∥ ≤ 1}
is weakly compact.
Theorem 2.13
Assume that X is a reflexive Banach space and let (xn ) be a bounded sequence in X. Then
there exists a subsequence (xnk ) that converges weakly.
Assume that X is Banach space such that every bounded sequence in X admits a weakly
convergent subsequence. Then X is reflexive.
95
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Propoistion 2.4
Assume that X is a reflexive Banach space and let Y ⊂ X be a closed linear subspace of X.
Then Y is reflexive.
Corollary 2.2
Propoistion 2.5
Example 2.3
96
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Chapter 3
Compact Operators
Definition 3.1
1. T is a compact operator.
Theorem 3.1
97
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Proof. Let T be a compact operator and suppose that T is not bounded. Then for each integer
n ≥ 1 there exists a unit vector xn such that ∥T xn ∥ ≥ n. Since the sequence (xn ) is bounded,
there exists a subsequence (T xnk ) which converges. But this contradicts ∥T xnk ∥ ≥ nk . Hence T
must be bounded. □
Theorem 3.2
T is compact ⇐⇒ for every bounded subset A ⊂ X, the set T (A) ⊂ Y is relatively compact.
Proof. (⇒): Suppose that T is compact. Let A ⊂ X be bounded and let (yn ) be sequence in
T (A). Then for each n ∈ N, there exists xn ∈ A such that ∥yn − T xn ∥ < 1n and the sequence (xn )
is bounded since A is bounded. Thus the sequence (T xn ) contains a convergent subsequence
because T is compact. Hence (yn ) contains a convergent subsequence with limit in T (A). This
shows that T (A) is compact.
(⇐): Suppose that for every bounded subset A ⊂ X the set T (A) ⊂ Y is relatively compact.
Then for any bounded sequence (xn ) in X the sequence (T xn ) lies in a compact set, and hence
contains a convergent subsequence. So T is compact. □
Theorem 3.3
T is compact ⇐⇒ for any sequence of vectors (xn ) in a closed unit ball B1 (0) ⊂ X, the
sequence (T xn ) has a convergent subsequence.
Theorem 3.4
Proof. 1. As T has finite rank, the space Im T is finite-dimensional normed space. More-
over, for any bounded sequence (xn ) in X, the sequence (T xn ) is bounded in Im T , hence
by the Bolzano-Weierstrass theorem this sequence must contain a covergent subsequence.
Thus T is compact.
98
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
2. If dim X is finite then the rank of T , r(T ) ≤ dim X, so r(T ) is finite, while if dim Y is finite
then the dimension of Im T ⊂ Y must be finite. Thus, in either case the result follows
from the first part.
□
Theorem 3.5
Let (T n )n≥1 be a sequence of compact linear operators from a normed space X into a Banach
spaces Y. If ∥T n − T ∥ → 0 then the limit operator T is compact.
Proof. We prove this theorem using a diagonal argument. Since T 1 is a compact operator, then
the sequence (T 1 (xn )) has a convergent (hence Cauchy) subsequence (T 1 (x1,m )) where x1,m is a
subsequence of the sequence (xn ). The subsequence (x1,m ) is bounded, so we can repeat the
argument with T 2 to produce a subsequence (x2,m ) of (x1,m ) with the property that (T 2 (x2,m ))
converges. We continue the process and then define a sequence (ym ) = (xm,m ). Note that (ym )
is a subsequence of (xn ) so it bounded, say by ∥ym ∥ ≤ c, and it has the property that for every
fixed n, the sequence (T n (ym )) is convergent and hence Cauchy.
We claim that (T (ym )) is a Cauchy sequence in Y. Let ϵ > 0. Since ∥T n − T ∥ → 0 there is
some l ∈ N such that ∥T l − T ∥ < 3cϵ . Moreover, since (T l (ym )) is Cauchy, there is some N > 0
such that ∥T l (y j ) − T l (yk )∥ < 3ϵ whenever j, k > N. Therefore, for j, k > N, we have
∥T (y j ) − T (yk )∥ ≤ ∥T (y j ) − T l (y j ) + T l (y j ) − T l (yk ) + T l (yk ) − T (yk )∥
≤ ∥T (y j ) − T l (y j )∥ + ∥T l (y j ) − T l (yk )∥ + ∥T l (yk ) − T (yk )∥
ϵ
< ∥T − T l ∥∥y j ∥ + + ∥T − T l ∥∥yk ∥
3
ϵ ϵ ϵ
< c + + c = ϵ,
3c 3 3c
this implies that (T (ym )) is Cauchy. Since Y is a Banach space, then (T (ym )) is convergent. Thus
we have proved that for any bounded sequence (xn ) ⊂ X, the sequence (T xn ) has a convergent
subsequence; hence T is compact. □
Example 3.1
Then T N is compact.
Indeed, for any bounded sequence (xn ) ⊂ l2 , the sequence (T N xn ) is a bounded sequence in
RN or CN ⊂ l2 . Since every bounded sequence in RN or CN has a convergent subsequence, it
follows that T N is compact.
99
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Example 3.2
Then T is compact.
For x ∈ l2 ,
X 1/2
∥T N x − T x∥2 = |λn xn |2
n>N
X 1/2
≤ sup |λn | |xn |2
n>N n>N
≤ sup |λn | ∥x∥2 .
n>N
Theorem 3.6
Proof. Since dim(X) = ∞, there exists a sequence of unit vectors (xn ) in X which does not
have any convergent subsequence. Hence the sequence (I xn ) = (xn ) cannot have a convergent
subsequence and thus the operator I is not compact. □
Corollary 3.1
Proof. Assume that T is invertible. Then the identity operator I = T −1 T on X must be compact.
But since X is infinite-diemensional this contradicts Theorem 6. □
We present further properties of compact operators. [For the proof see [1]]
100
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Theorem 3.7
The converse is not true, in general, when Y is Banach space, but it holds when Y is a Hilbert
space.
Theorem 3.8
Example 3.3
Example 3.4
Consider a bounded sequence (en ), where en = (0, · · · , 0, 1, 0, · · · ) with the n th position equals
to 1. Thus for any n , m, √
∥Ien − Iem ∥ = ∥en − em ∥ = 2.
This implies that any subsequence of (Ien ) cannot be Cauchy and hence cannot converge. So I
is not compact.
Example 3.5
K : [0, 1] × [0, 1] → R
(x, y) 7→ K(x, y)
be continuous. Define T : X → X by
Z 1
T ( f )(x) = K(x, y) f (y)dy.
0
101
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Then T is compact.
Hence T is bounded.
To prove that T is compact, let A ⊂ X be bounded. Let us prove that T (A) is relatively compact.
First we prove that T (A) is equicontinuous. Let ϵ > 0. Since K is continuous on comapct
metric space then K is uniformly continuous so ∃δ > 0 such that ∀x1 , x2 ∈ [0, 1], |x1 − x2 | < δ
implies |K(x1 , y) − K(x2 , y)| < ϵ. For any f ∈ A,
Z 1
|T ( f )(x1 ) − T ( f )(x2 )| = (K(x1 , y) − K(x2 , y)) f (y)dy
0
Z 1
≤ ∥ f ∥∞ |K(x1 , y) − K(x2 , y)|dy
0
≤ ∥ f ∥∞ ϵ ≤ sup ∥ f ∥∞ ϵ < (1 + sup ∥ f ∥∞ )ϵ.
f ∈A f ∈A
Thus T (A) is equicontinuous. For any x ∈ [0, 1], {T ( f )(x); f ∈ A} is bounded in R so it is rela-
tively compact. Hence Arzéla-Ascoli theorem implies that T (A) is relatively compact.
Definition 3.2
Let (xn ), x ∈ X, we say that (xn ) converges to x weakly if for any φ ∈ X ∗ we have that
w
limn→∞ φ(xn ) = φ(x). The point x is called the weak limit of (xn ), written xn → x or
xn ⇀ x.
Remark 3.1
In particular, if X = (H, ⟨., .⟩) is a Hilbert space then xn ⇀ x if and only if ⟨y, xn ⟩ →
⟨y, x⟩, ∀y ∈ H.
As a consequence of Theorem Kakutani 12, in a reflexive space any bounded sequence has a
weakly convergent subsequence.
102
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Properties:
xn → x then xn ⇀ x.
Theorem 3.9
Let T be a compact operator from a normed space X to a normed space Y. Then T maps
any weakly convergent sequence into a strongly convergent sequence (a sequence that
converges in the norm of X).
since φT ∈ X ∗ . Now suppose that T xn does not converge in the norm to T x0 . Then there exists
some ϵ > 0 and a subsequence xnk such that ∥T xnk − T x0 ∥ ≥ ϵ. Since xnk ⇀ x, we know that
(xnk ) is bounded. By the compactness of T , (T xnk ) has a convergent subsequence (T xn′k ) such
that T xn′k → ȳ ∈ Y. By the uniqueness of the weak limit, it follows that ȳ = T x0 , but this
contradicts the fact that ∥T xnk − T x0 ∥ ≥ ϵ. □
Remark 3.2
xnk → x in L2 (Ω).
103
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
That is, weak convergence in H 1 (Ω) implies the existence of a subsequence that is strongly
convergent in L2 (Ω).
Theorem 3.10
Let Ω ⊂ Rn be a bounded domain with boundary ∂Ω partitioned as ∂Ω = Γ∪Γ′ and |Γ| > 0.
Let H = u ∈ H 1 (Ω) | u|Γ = 0 . Then there exists a constant C(Ω) such that
Z Z
2
u dx ≤ C(Ω) |∇u|2 dx, ∀u ∈ H.
Ω Ω
Definition 3.3
Let H be a Hilbert space and let T ∈ L(H). The resolvent of T , denoted by ρ(T ), is the set
of complex numbers λ such that (T − λI) : H → H is one-to-one and onto. The spectrum
of T , denoted by σ(T ), is the complement of the resolvent in C,
σ(T ) = C \ ρ(T ).
104
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
Introduction to Functional Analysis MEAS Len
Remark 3.3
If T − λI is one-to-one and onto then the open mapping theorem implies that (T − λI)−1 is
bounded. Hence, when λ ∈ ρ(T ) both T − λI and (T − λI)−1 are bounded.
Now we introduce the adjoint operators, which is guaranteed to exist by Riesz representation
theorem 33.
Theorem 3.11
Let T ∈ L(H) be a bounded operator on H. Then there exists a unique operator T ∗ ∈ L(H)
such that
⟨T x, y⟩ = ⟨x, T ∗ y⟩ for all x, y ∈ H.
The operator T ∗ is called the adjoint of T .
Let H be a Hilbert space and let T : H → H be a compact self-adjoint operator. Then there
exists an orthornormal set (en )n≥1 consisting of eigenvectors of T with eigenvalues λn such
that ∞
X
Tx = λn ⟨x, en ⟩en , x ∈ H. (3.2.1)
n=1
To prove this theorem we need the following Lemmas. For the details proof see [10].
105
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
MEAS Len Introduction to Functional Analysis
Lemma 3.1
Lemma 3.2
• for each µ > 0 the space spanned by the eigenvectors corresponding to all λk with
|λk | > µ is finite dimensional.
The expression for the norm of self-adjoint operator is ∥T ∥ = sup∥x∥=1 |⟨x, T x⟩|. Hence we either
have
∥T ∥ = sup |⟨x, T x⟩| or ∥T ∥ = − inf |⟨x, T x⟩|.
∥x∥=1 ∥x∥=1
Lemma 3.3
Therefore, the representation (3.2.1) of the compact self-adjoint operator T is an infinite dimen-
sional version of the well-known result in finite dimensional linear algebra that a self-adjoint
matrix can be diagonalized by a basis consisting of eigenvectors of the matrix.
106
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44
References
[1] B. P. Rynne, M. A. Youngson, Linear Functional Analysis, 2nd edition, Springer (2008).
[3] H. Brezis, Functional Analysis, Sobolev Spaces and Partial Differential Equations,
Springer (2011).
[5] J. Muscat, Functional Analysis: An Introduction to Metric Spaces, Hilbert Spaces, and
Banach Algebras, Springer (2014).
[7] M. Einsiedler, T. Ward, Functional Analysis, Spectral Theory, and Applications, Springer
(2017).
[9] S. Salsa, Partial Differential Equations in Action: From Modelling to Theory, Springer
(2016).
[11] W. Rudin, Principles of Mathematical Analysis, McGraw Hill, 3rd edition (1976).
107
AMS Open Math Notes: Works in Progress; Reference # OMN:202208.111337; Last Revised: 2022-08-28 11:06:44