Functional Analysis
Functional Analysis
Course-MMDSC 3.3
Functional Analysis
Programme Name: M.Sc. Mathematics (CBCS) Year/Semester: III Semester COURSE WRITER
Course Code: MMDSC 3.3 Course Name: Functional Analysis
Credit: 4 Unit Number : 1-16 Prof. H. S. Ramane Block 3.3 A to Block 3.3 D
Professor (Block I – IV)
COURSE DESIGN COMMITTEE
Department of Studies in Mathematics (Unit 01 to Unit 16)
Dr. Sharanappa. V. Halse Chairman Karnatak Unuiversity, Dharwad
Vice Chancellor
Karnataka State Open University
Mukthagangothri, Mysuru-570006
Prof. N. Lakshmi Member
COURSE EDITOR
Dean (Academic)
Karnataka State Open University
Mukthagangothri, Mysuru-570006 Dr. K. R. Vasuki Block 3.3 A to Block 3.3 D
Dr. Pavithra. M Course coordinator Professor (Block I – IV)
Assistant Professor DOS in Mathematics (Unit 01 to Unit 16)
DoS in Mathematics, KSOU, Mukthagangothri, Mysuru-06 University of Mysore, Mysuru
EDITORIAL COMMITTEE
The Registrar
1. Dr. K. Shivashankara Chairman
Karnataka State Open University
BOS Chairman(PG), DoS in Mathematics, KSOU
Mukthagangothri, Mysuru-570006
Associate Professor, Yuvaraja College,
University of Mysore, Mysuru-06
Developed by the Department of Studies in Mathematics under the guidance of Dean
2. Mr. S. V. Niranjana Member & Convener
(Academic), KSOU, Mysuru.
Coordinator, (DoS in Mathematics)
Karnataka State Open University, 2023.
Assistant Professor, DoS in Physics,
All rights reserved. No part of this work may be reproduced in any form or any other means, without
KSOU, Mysuru-06
permission in writing from the Karnataka State Open University.
3. Dr. Pavithra. M Member Further information on the Karnataka State Open University Programmes may be obtained from the
Assistant Professor University’s Office at Mukthagangothri, Mysuru – 570 006.
DoS in Mathematics, KSOU, Mysuru-06
Dr. H. S. Ramane
Department of Mathematics, Karnatak University, Dharwad - 580003
i
3.3 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.4 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
ii iii
7.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 11.4 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
7.4 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
12 Unit 12: Banach - Steinhaus Principle of Uniform Boundedness 127
10 Unit 10: Open Mapping Theorem 109 14 Unit 14: Properties of Hilbert Spaces 142
iv v
16 Unit 16: Riesz Lemma and Riesz Representation Theorem 156 Block - I
16.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
vi 1
Remark 1.2.2. (1) A metric d is always non-negative. For x, y ∈ X
it follows that
d(x, y) + d(y, x) ≥ d(x, x)
2 3
Definition 1.2.7. Let (X, d) be a metric space. Let A ⊆ X. A point If such an isometry exists then (X, d1 ) and (Y, d2 ) are said to be
x ∈ X is said to be a limit point of A if every neighbourhood (open isometric to each other.
ball around x) of x contains atleast one point of A.
Theorem 1.2.15. A set G is open in a metric space (X, d) if and
Definition 1.2.8. Let (X, d) be a metric space. Let A ⊆ X, then only if G is a union of open spheres.
closure of A denoted by A is nothing but set of all points x ∈ X such
Proof. Suppose G is open set in a metric space (X, d).
that either x ∈ A or x is a limit point of A.
x ∈ G ⇒ ∃ rx > 0 ∋ x ∈ Srx (x) ⊆ G.
Definition 1.2.9. Let (X, d) be a metric space. A set F ⊆ X is said
to be closed if and only if F = F (F contains all its limit points).
⇒ {x} ⊆ Srx (x) ⊆ G.
Definition 1.2.10. Let (X,d) be a metric space. Let A ⊆ X is said [ [
⇒ {x} ⊆ Srx (x) ⊆ G.
to be dense in X or everywhere dense in X if and only if A = X. x∈G
[
x∈G
⇒ G⊆ Srx (x) ⊆ G.
Definition 1.2.11. Let A ⊆ X. Then diameter of A is given by x∈G
[
⇒ G= Srx (x).
d(A) = sup d(x, y), (d(A) ≤ ∞). x∈G
x,y∈A
S
Definition 1.2.12. A set A ⊆ X is said to be a bounded set if d(A) < Suppose G = Sα (arbitrary union of open sphere i.e., ∀ α, Sα
α∈λ
∞ (d(A) ̸= ∞). is an open sphere in G).
S
If x ∈ G ⇒ ∃ α ∈ λ ∋ x ∈ Sα and Sα ⊆ Sα = G.
Definition 1.2.13. Let A, B be the subsets of X. Then distance α∈λ
4 5
1 ≤ i ≤ n or d(x, ai ) > 0, 1 ≤ i ≤ n. Set r = min d(r, ai ) ⇒ x is not a limit point of F .
1≤i≤n
/ A or y ∈ A′ ⇒ y ∈ Sr (x) ⊆ A′ .
Therefore y ̸= ai or y ∈ Solution: Let {x ∈ X : d(x, A) = 0} = B.
Theorem 1.2.17. A set F is closed set in (X, d) if and only if F ′ is If x ∈ A ⇒ either x ∈ A or x is a limit point of A. If x ∈ A,
Proof. Suppose F is a closed set in (X, d). Suppose x is a limit point of A, ⇒ for every r > 0, Sr (x) ∩ A ̸= ∅.
Let x ∈ F ′ , then x ∈
/ F. ∴ y ∈ Sr (x) ∩ A ⇒ y ∈ A and d(x, A) < r, ∀ r > 0.
⇒ x ∈ Sr (x) ⊆ F ′ . If x ∈ B ⇒ d(x, A) = 0.
6 7
⇒ x ∈ A or B ⊆ A. Cauchy sequence if given ϵ > 0, ∃ a positive integer N such that
Hence A = B. m, n ≥ N
⇒ |xm − xn | < ϵ or dR (xm , xn ) < ϵ.
(2) Prove that a set A in X is dense in X or A = X, if and only if the Definition 1.3.2. A sequence {xn } of real numbers is said to be a
1.3 Complete Metric Space Definition 1.3.6. A metric space (X, d) is said to be a Complete
metric space if every Cauchy sequence is convergent in X.
Simplest example of complete metric space is R or real line. The
following three fundamental concepts of real analysis are to describe Example 1.3.7. (1) X = Rn and
the completeness of R1 .
p
d((x1 , x2 , . . . , xn ), (y1 , y2 , . . . , yn )) = (x1 − y1 )2 + (x2 − y2 )2 + · · · + (xn − yn )2 .
Definition 1.3.1. A sequence {xn } of real numbers is said to be a Then (X, d) is a metric space.
8 9
(n) (n) (n) bers are non-negative).
Put x = (x1 , x2 , . . . , xn ), {x(n) } = {(x1 , x2 , . . . , xn )} is a se-
quence in Rn . d(z, w) = 0 ⇐⇒ |zi − wi |2 = 0, ∀ 1 ≤ i ≤ n.
Suppose {x(n) } is a Cauchy sequence ⇐⇒ |zi − wi | = 0, ∀ 1 ≤ i ≤ n.
(i) d(z, w) ≥ 0 (sum of sequences of absolute values of complex num- d(z, w) ≤ d(z, t) + d(t, w).
10 11
∴ (Cn , d) is metric space. Definition 1.3.9. Let (N, +, ·, F ) be a linear space where F = R or
(m) (m) (m)
Suppose {Z (m) } = {z1 , z2 , . . . , zn } is Cauchy sequence. C. Then a Norm on N is || · || : N → R+ satisfying the following
(m) properties.
⇒ {zi } is Cauchy sequence in C and C is complete.
(i) ||x|| ≥ 0 and ||x|| = 0 ⇐⇒ x = 0N .
⇒ {z (m) } is also convergent.
(ii) ||λx|| = |λ|||x||.
n
Hence (C , d) is a Complete metric space. (iii) ||x + y|| ≤ ||x|| + ||y||, ∀ x, y ∈ N and λ ∈ F .
Definition 1.3.8. (N, +, ·, F ) is said to be a Linear space over a field Definition 1.3.10. If (N, +, ·, F ) is a linear space over F = R or C
F if N ̸= ∅ and if the following axioms are satisfied. and if || · || : N → F is a norm on N , then (N, || · ||) is called Normed
linear space.
(i) x, y ∈ N ⇒ x + y ∈ N (Vector addition).
(viii) α, β ∈ F, x ∈ N ⇒ (α + β) · x = α · x + β · x.
(ix) α, β ∈ F, x ∈ N ⇒ α · (β · x) = (αβ) · x.
12 13
(ii) Proof. (i) Consider
(iii) stant).
14 15
(a) |f (x)| ≥ 0 ⇒ ||f || ≥ 0
∴ (B ∗ (X), || · ||) is a normed linear space. Lemma 1.3.14. Suppose (X, d) is a metric space and {xn } is a con-
(iii) Suppose {fn } is a Cauchy sequence in B ∗ (X). vergent sequence with infinitely many distinct points. Then the limit
⇒ For every ϵ > 0, ∃ a positive number N , such that m, n ≥ N , of {xn } is same as limit point of S = {x1 , x2 , . . . , xn , . . .}.
16 17
Proof. Suppose if possible x = lim xn is not a limit point of S.
n→∞
18 19
Definition 1.3.17. A function f : X → F is said to be continuous Choose δ = min(δ1 , δ2 ). Then
on X, if it is continuous for all x ∈ X. d(x, x0 ) < δ ⇒ |(f + g)(x) − (f + g)(x0 )|
Note 1.3.18. Isometry between two metric space is obviously contin- = |{f (x) + g(x)} − {f (x0 ) + g(x0 )}|
uous. = |{f (x) − f (x0 )} + {g(x) − g(x0 )}|
Theorem 1.3.19. Let C ∗ (X) = {f : X → F | (X, d) is a metric ≤ |f (x) − f (x0 )| + |g(x) − g(x0 )|
ϵ ϵ
space, F = R or C} < +
2 2
where f is a continuous and bounded on X. Then C ∗ (X) ⊂ B ∗ (X) = ϵ
and further
∴ f + g is continuous on X.
∗ ∗
(i) C (X) is a subspace of B (X).
|f (x)| ≤ kf , |g(x)| ≤ kg
(ii) C ∗ (X) is a closed subspace of B ∗ (X) as a consequence, C ∗ (X) is
a complete metric space. ⇒ |(f + g)(x)| = |f (x) + g(x)| ≤ |f (x)| + |g(x)|
≤ kf + kg
Proof. (i) Suppose f, g ∈ C ∗ (X) and α, β ∈ F .
= kf +g
∗
To show that: (a) f + g ∈ C (X)
∴ f + g is bounded on X.
(b) αf ∈ C ∗ (X)
Hence f + g ∈ C ∗ (X).
⇐⇒ αf + βg ∈ C ∗ (X).
ϵ
Next, consider αf . Since f is continuous at x0 , for > 0.
Consider (f + g)(x) and x0 ∈ X. Since f, g ∈ C ∗ (X), given ϵ > 0, |α|
∃ δ1 > 0, δ2 > 0 such that ∃ δ = δ(ϵ, |α|) such that d(x, x0 ) < δ
ϵ
⇒ |f (x) + f (x0 )| <
|α|
ϵ ϵ
d(x, x0 ) < δ1 ⇒ |f (x) − f (x0 )| < ⇒ |α||f (x) + f (x0 )| < |α| =ϵ
2 |α|
ϵ
d(x, x0 ) < δ2 ⇒ |g(x) − g(x0 )| < . ⇒ |(αf )(x) − (αf )(x0 )| < ϵ
2
∴ αf is continuous on X.
20 21
|f (x)| ≤ kf , ⇒ |(αf )(x)| = |α||f (x)| ≤ |α|kf = kαf ∴ C ∗ (X) is a closed subspace of B ∗ (X).
Consider,
⇒ |{f (x) − f0 (x)} + {f0 (x) − f0 (x0)} + {f0 (x0 ) − f (x0 )}|
To show that f (x) is bounded on X: Theorem 1.3.22 (Metric completion is unique upto isome-
try:). All completion of a given metric space are isometric.
|f (x)(ξ)| = |d(x, ξ) − d(ξ, z)|
= |d(x, ξ) − d(z, ξ)| Proof. Suppose (X ∗ , d∗ ) and (X ∗∗ , d∗∗ ) are any two metric completions
≤ d(x, z) = kf (x) < ∞, ∀ ξ ∈ X. of a given metric space (X, d). Then by Theorem 1.3.20, ∃ f, g : X →
C ∗ (X) two isometries such that X ∗ = f (X) and X ∗∗ = g(X).
∴ f (x) is bounded on X and hence f (x) ∈ C ∗ (X) is well defined.
= f (X) ∼
Without loss of generality, X(∼ = g(X)) itself can be taken to
To show that f (x) is an isometry: be dense in both X ∗ and X ∗∗ .
Consider,
Claim: ∃ F : X ∗ → X ∗∗ such that d∗∗ (F (x∗ ), F (y ∗ )) ≤ d∗ (x∗ , y ∗ ).
d∗ (f (x), f (y)) = sup |f (x)(ξ) − f (y)(ξ)| Let x∗ ∈ X ∗ . Since X is dense in X ∗ or X = X ∗ , x∗ is a limit point
ξ∈X
= sup |{d(x, ξ) − d(ξ, z)} − {d(y, ξ) − d(ξ, z)}| of X. ∃ a sequence {xn } in X such that lim xn = x∗ . So, {xn } is a
ξ∈X n→∞
= sup |{d(x, ξ) − d(y, ξ)}| convergent sequence in X. Hence {xn } is a Cauchy sequence in X.
ξ∈X
≤ d(x, y). Then, X ⊆ X = X ∗∗ ⇒ {xn } is a Cauchy sequence in X ∗∗ and
24 25
X ∗∗ is Complete. So, ∃ x∗∗ ∈ X ∗∗ such that lim xn = x∗∗ . 1.4 Exercises
n→∞
x∗ → x∗∗ will be well defined provided {xn } is unique.
To prove that {xn } is unique: (1) Prove that a set A in X is dense in X or A = X.
Suppose, if possible there is another sequence {x′n } in X converging (i) ⇐⇒ the only closed superset of A is X.
⇒ xn = x′n .
1.5 References
Hence, for every x∗ ∈ X ∗ , there is well defined element x∗∗ ∈ X ∗∗ .
1. G. F. Simmons - Introduction to Topology and Modern Analysis,
Define F : X ∗ → X ∗∗ by F (x∗ ) = x∗∗ .
Tata McGraw-Hill, New Delhi, 2004.
Suppose x∗ , y ∗ ∈ X ∗ . Then ∃ {xn } and {yn } in X such that lim xn =
n→∞
x∗ , lim yn = y ∗ . 2. A. E. Taylor - Introduction to Functional Analysis, Willey, New
n→∞
Since X = X ∗∗ , lim xn = x∗∗ , lim yn = y ∗∗ . York, 1958.
n→∞ n→∞
Consider,
−−−−−−−−−−−−−−−−−−−−−
= lim d(xn , yn )
n→∞
= d∗ (x∗ , y ∗ ).
26 27
Definition 2.2.2. Let (X, d) be a metric space and let T : X → X be
a mapping from X into X. Then a point x ∈ X is said to be a fixed
point of T , if Tx = x.
In this unit, we introduce the concept of Banach Contraction Mapping Proof. : Let x0 ∈ X. Define sequence {xn } in X as follows:
xn = Txn−1 = Txn0 .
Definition 2.2.1. Let (X, d) be a metric space and let T : X → X be a
mapping from X into X. Then T is said to be a contraction mapping, To show that {xn } is a Cauchy sequence:
1
if there exists α ∈ R such that 0 < α < 1 and d(Tx , Ty ) ≤ αd(x, y),
∀ x, y ∈ X.
28 29
Let us observe that, Consider,
d(x0 , x1 )
d(x0 , xn ) = .
1−α
30 31
x x
dy
Z Z
It follows that d(x0 , xn ) < r, (∵ d(x0 , Tx0 ) < (1 − α)r).
⇐⇒ dx = f (t, y(t))dt
x0 dx
This shows that all xn ’s are in B r (x0 ). Since xn → x and B r (x0 ) is Zx0x
or y(x) − y(x0 ) =
f (t, y(t))dt
closed, it follows that x ∈ B r (x0 ). The assertion of the theorem now x0
Z x
follows from the proof of the Theorem 2.2.4. or y(x) = y0 + f (t, y(t))dt.
x
Rx 0
Define T : X → X by Ty(x) = y0 + x0 f (t, y(t))dt, where X is the
2.3 Applications space of all solutions of the initial value problem.
has a unique solution on [x0 − α, x0 + α]. (x, y) ∈ Sδ ((x0 , y0 )) ⇒ |f (x, y) − f (x0 , y0 )| < 1.
32 33
Then R can hold all the solutions passing through (x0 , y0 ). Next, T is continuous:
Z x1 Z x2
Define, X = {ϕ : [x0 − α, x0 + α] → [y0 − αk, y0 + αk]}, where ϕ is |Tϕ(x1 ) − Tϕ(x2 ) | = y0 + f (t, ϕ(t))dt − y0 + f (t, ϕ(t))dt
continuous and ϕ(x0 ) = y0 . Z x2 x0 x0
= − f (t, ϕ(t))dt
To show that X is complete:
Z x2 x1
Clearly, X ⊆ C ∗ ([x0 −α, x0 +α]) = {ϕ : [x0 −α, x0 +α] → R1 | ϕ is continuous} ≤ |f (t, ϕ(t))|dt
x1
C ∗ ([x0 −α, x0 +α]) is a complete metric space, ψ ∈ C ∗ ([x0 −α, x0 +α]) ≤ k|x1 − x2 |.
and ψ ∈ X
Given ϵ > 0, choose δ = kϵ , then continuity follows.
⇒ ∃ {ϕn } in X such that ψ(x) = lim ϕn (x), ∀ x ∈ [x0 − α, x0 + α].
n→∞ Rx
Finally, to show that T : X → X, by Ty = y0 + x0 f (t, y(t))dt is
For all n, y0 − αk ≤ ϕn (x) ≤ y0 + αk.
a contraction mapping:
⇒ y0 − αk ≤ ψ(x) ≤ y0 + αk.
For every, ϕ1 , ϕ2 ∈ X, consider,
ψ(x0 ) = lim ϕn (x0 ) = lim y0 = y0 .
n→∞ n→∞
∴ ψ ∈ X or X is a closed subspace of C ∗ ([x0 − α, x0 + α]). d∗ (Tϕ1 , Tϕ2 ) = sup |Tϕ1 (x) , Tϕ2 (x) |
x∈[x0 −α,x0 +α]=I
Z x Z x
Hence, using the standard theorem, X is a complete metric space.
= sup y0 + f (t, ϕ1 (t))dt − y0 + f (t, ϕ2 (t))dt
Rx x∈I x0 x0
To show that T : X → X, by Ty = y0 + x0 f (t, y(t))dt is a Z x
= sup {f (t, ϕ1 (t)) − f (t, ϕ2 (t))}dt
contraction mapping. x∈I x0
Z x
First, let us show that if ϕ(x) ∈ [y0 − αk, y0 + αk], ≤ sup |f (t, ϕ1 (t)) − f (t, ϕ2 (t))|dt
x∈I x0
then Tϕ(x) ∈ [y0 − αk, y0 + αk]. Z x
Z x ≤ sup M |ϕ1 (t) − ϕ2 (t)|dt
x∈I x0
|Tϕ(x) − y0 | = f (t, ϕ(t))dt
≤ sup M |ϕ1 (t) − ϕ2 (t)||x − x0 |
Z xx0 x∈I
≤ |f (t, ϕ(t))|dt ≤ d∗ (ϕ1 (t), ϕ2 (t))M α
x0
Z x
< kdt ∴ d∗ (Tϕ1 , Tϕ2 ) ≤ d∗ (ϕ1 (t), ϕ2 (t)), (∵ M α < 1).
x0
= k(x − x0 ) ≤ kα, (∵ x ∈ [x0 − α, x0 + α] ⇒ (x − x0 ) < α). Let us apply Banach Contraction Mapping Theorem for
Z x
Tϕ(x) = y0 + f (t, ϕ(t))dt.
⇒ −kα ≤ Tϕ(x) − y0 ≤ kα, or Tϕ(x) ∈ [y0 − αk, y0 + αk]. x0
34 35
There is a unique y(x) ∈ X such that Ty(x) = y(x) or 2. A. E. Taylor - Introduction to Functional Analysis, Willey, New
Z x
York, 1958.
y(x) = y0 + f (t, y(t))dt,
x0
dy
= f (x, y), y(x0 ) = y0
dx
2.4 Exercises
2.5 References
36 37
Definition 3.1.3. Let (X, d) be a metric space. A subset Y ⊆ X is
said to be everywhere dense, if Y = X.
OR Int(Y ) = X.
38 39
Theorem 3.1.8 (Baire’s Category Theorem). Every nonempty Sr0 (x0 ) ⊇ Sr1 (x1 ) ⊇ Sr2 (x2 ) ⊇ · · · ⊇ Srn (xn ) ⊇ · · · .
complete metric space is of second category. Claim: {xn } is a Cauchy sequence in X.
1
xm ∈ Srn (xn ) ⇒ d(xm , xn ) < rn < n
Proof. we need following Lemma.
1
xn ∈ Srm (xm ) ⇒ d(xm , xn ) < rm < m
Lemma: Suppose {Gn : n = 1, 2, . . .} is a sequence of everywhere 1
∞
T ⇒ d(xm , xn ) < min(m,n) → 0 as m, n → ∞.
dense subsets of a complete metric space (X, d). Then Gn is also
n=1 ∴ {xn } is a Cauchy sequence in X.
everywhere dense in X.
Since, X is complete. We have lim xn = x ∈ X.
n→∞
Proof of the Lemma: We know that Y is everywhere dense in X. ⇒ x ∈ Srn (xn ), ∀ n ≥ n0 .
⇐⇒ Y intersects every open sphere in X. ⇒ x ∈ Srn (xn ), ∀ n
i.e. x ∈ X and r > 0 ⇒ Y ∩ Sr (x) ̸= ∅. (∵ Sr0 (x0 ) ⊇ Sr1 (x1 ) ⊇ Sr2 (x2 ) ⊇ · · · ⊇ Srn (xn ) ⊇ · · · )
∞
T
It is enough, we show that A = Gn intersects any open sphere in
n=1 rn (xn ) ⊆Gn , x ∈ Gn , ∀ n
since, S
∞
X. ⇒x∈
T
Gn = A and x ∈ Sr0 (x0 )
n=1
Let x0 ∈ X and r0 > 0 be arbitrary. Then it is enough to show that ⇒ Sr0 (x0 ) ∩ A ̸= ∅.
∞
A ∩ Sr0 (x0 ) ̸= ∅. ∴A=
T
Gn everywhere dense in X.
n=1
Given, G1 is everywhere dense in X.
Proof of the Theorem 3.1.8:
⇒ Sr0 (x0 ) ∩ G1 ̸= ∅, we can find x1 ∈ Sr0 (x0 ) ∩ G1 , choose, 0 < r1 < 1.
Suppose {Yn : n = 1, 2, . . .} is a sequence of nowhere dense sub-
Then Sr1 (x1 ) ⊆ Sr0 (x0 ) ∩ G1 and Sr0 (x0 ) ⊇ Sr1 (x1 ).
sets of X.
Next, G2 is also given to be everywhere dense in X.
∞
S
⇒ Sr1 (x1 ) ∩ G2 ̸= ∅. So, we can find x2 ∈ Sr1 (x1 ) ∩ G2 , choose, 0 < Claim: X ̸= Yn or ∃ x ∈ X such that x ∈
/ Yn , ∀ n.
n=1
r2 < 12 . Then Sr2 (x2 ) ⊆ Sr1 (x1 ) ∩ G2 and Sr0 (x0 ) ⊇ Sr1 (x1 ) ⊇ Sr2 (x2 ).
40 41
S ∞
∅ = Int(Yn ) = Gn ( where Gn is an open subset of Yn ), Proof. First, we shall show that
T
Fn cannot have more than one
Gn ⊆Yn n=1
42 43
Since, F1 ⊇ F2 ⊇ · · · ⊇ FN −1 ⊇ FN , x ∈ Fn , n = 1, 2, . . . , N − 1 ⇒ Choose, xn ∈ S1/2n−1 (xn−1 ) and rn = 1/2n−1 , so that S1/2n−1 (xn ) ∩
∞
T ∞
T
∴ x ∈ Fn , ∀ n or x ∈ Fn or Fn ̸= ∅. An = ∅.
n=1 n=1
Case II: Suppose S = {x1 , x2 , . . . , xn , . . .} is infinite. Put Fn = S1/2n (xn ) so that F1 ⊇ F2 ⊇ · · · ⊇ Fn and Fn ∩ An = ∅.
Since {xn } converges to x, x is a limit point of S. Also, d(Fn ) = diameter of the closed sphere with radius 1/2n =
⇒ x is also a limit point of Fn and Fn is a closed set. Hence by Cantor’s Intersection Theorem,
∞
∞ ∞ T
⇒ x ∈ Fn , for n = 1, 2, . . .. ∴ x ∈
T
Fn or
T
Fn ̸= ∅ Fn = {x} for some x ∈ X.
n=1
n=1 n=1
∞
T ⇒ x ∈ Fn , ∀ n = 1, 2, . . .
Hence Fn has exactly one point.
n=1 But Fn ∩ An = ∅.
Problem 3.1.10. Use of Cantor’s Intersection Theorem to prove Baire’s ∴x∈
/ An , ∀ n = 1, 2, . . ..
Category Theorem.
We may continue this process for every n = 1, 2, . . ., so that we may is dense, but of the first category in S, where S denotes the set of all
44 45
3.3 References
−−−−−−−−−−−−−−−−−−−−−
Unit 4: Ascoli-Arzela Theorem
4.1 Introduction
46 47
Definition 4.2.4. Let (X, d) be a metric space. (X, d) is said to be ⇒ S = {x1 , x2 , . . . , xn , . . .} will be an infinite set in X.
a sequentially compact, if every sequence {xn } in X has a convergent Since, X has Balzano-Weierstrass property, A has a limit point x ∈ X.
subsequence {xnk }. ⇒ For every k > 0, ∃ xnk ∈ S such that xnk ∈ Srnk (x) ∩ S(̸= ∅).
⇒ {xnk } is a convergent subsequence of {xn }.
Definition 4.2.5. Let (X, d) be a metric space. (X, d) is said to have
∴ X is also sequentially compact.
Balzano - Weierstrass property, if every infinite subset A ⊆ X has a
limit point. Theorem 4.2.7. Every compact metric space has Balzano-Weierstrass
Theorem 4.2.6. Let (X, d) be a metric space. Then X is sequentially property. Hence it is also sequentially compact.
Proof. Suppose X is sequentially compact and suppose A is any infi- infinite subset of X. Suppose if possible A has no limit point in X.
⇒ {xn } is a sequence in X and X is sequentially compact. Then {Gi } is a collection of open subsets of X. We may make {Gi }
Case I: Suppose {xn } = {x1 , x2 , . . . , xN −1 , x, xN +1 , . . . , xN +k , x, . . .}. totally bounded, if for every ϵ > 0, ∃ A = {a1 , a2 , . . . , aN } ⊆ X such
48 49
Definition 4.2.9. Let Y be a subset of a metric space (X, d). Then Y So, for {xn } there is a convergent subsequence {xnk } in X with
is said to be totally bounded, if for every ϵ > 0, ∃ A = {a1 , a2 , . . . , aN } ⊂ lim xnk = x.
nk →∞
X such that
N
[ 0 ≤ d(xn , x) ≤ d(xn , xnk ) + d(xnk , x) → 0 as n, nk → ∞.
Y ⊆ Sϵ (ak ).
k=1
∴ {xn } converges to x in X.
Note 4.2.10. (1) A is also called as ϵ-net of X.
Hence X is complete.
(2) x ∈ X ⇒ ∃ at least one ak ∈ A such that d(x, ak ) < ϵ.
Next, we shall show that X is also totally bounded.
Observation 4.2.11. (1) Suppose Y1 ⊆ Y2 ⊆ X and (X, d) is a metric
Suppose if possible, X is not totally bounded.
space. Then Y2 is totally bounded ⇒ Y1 is also totally bounded
N
S ∃ ϵ > 0, such that X has no ϵ-net or X has no covering by finitely
(∵ Y1 ⊆ Y2 ⊆ Sϵ (ak )).
k=1 many open spheres with radii ϵ. We shall construct a sequence {xn }
(2) A subset Y of a metric space (X, d) is totally bounded if and only
such that it has no convergent subsequence. This means X is not
if Y is totally bounded.
sequentially compact which will be a contradiction.
(3) A subset Y in a metric space (X, d) is totally bounded ⇒ Y is also
Let x1 ∈ X be any point in X. Since, X ̸= Sϵ (x)
bounded.
∃ x2 ∈ X \ Sϵ (x1 ) (otherwise, X = Sϵ (x1 ) ⇒ X is totally bounded
(4) A bounded set in a metric space need not be totally bounded.
which is against to the contra assumption).
(5) In Rn , every bounded set is also totally bounded.
⇒ d(x1 , x2 ) ≥ ϵ.
Theorem 4.2.12. Every sequentially compact metric space is both
Suppose, we have constructed x1 , x2 , . . . , xn ∈ X such that d(xi , xj ) ≥
complete and totally bounded.
ϵ, j = 1, 2, . . . , n.
n
S
Proof. First, we shall show that a sequentially compact metric space Since, X is not totally bounded, X ̸= Sϵ (xk )
k=1
n
(X, d) is complete.
S
⇒ ∃ xn+1 ∈ X \ Sϵ (xk ) and hence d(xn+1 , xi ) ≥ ϵ, ∀ i = 1, 2, . . . , n.
k=1
Let {xn } be any Cauchy sequence in X. By definition, d(xm , xn ) → Thus, inductively we have constructed a sequence {xn } in X such that
nition, every sequence in X has a convergent subsequence. ⇒ {xn } can not have a Cauchy sequence, ({xn } is convergent ⇒ {xn }
50 51
is Cauchy sequence if and only if {xn } is not Cauchy sequence ⇒ {xn } Suppose we have constructed,
is not convergent).
S1 = S(x1 ; 1/2),
⇒ {xn } can not have a convergent subsequence.
S2 = S(x2 ; 1/22 ),
Theorem 4.2.13. Let (X, d) be a complete and totally bounded metric ..
.
space. Then X is compact.
Sn = S(xn ; 1/2n ),
52 53
Next, ∴ For large value of k, Sk ⊆ S(x; ϵ) ⊆ G ∈ G
Sk is covered by a single member of G.
d(xm , xn ) ≤ d(xm , xm+1 ) + d(xm+1 , xm+2 ) + · · · + d(xn−1 , xn )
1 1 1 This is a contradiction.
< m−1 + m + · · · + n−2
2 2 2
Theorem 4.2.14. Let (X, d) be a metric space. Then the following
1 1 1
= m−1 1 + + · · · + n−m−1
2 2 2 are equivalent:
1 − rn+1
∵ GP = 1 + r + r2 + · · · + rn = (1) X is compact.
1−r
1 (1 − 1/2n−m ) (2) X is sequentially compact.
= m−1
2 (1 − 1/2)
(3) X is complete and totally bounded.
1 1
= 2 m−1 1 − n−m
2 2
1
1
Proof. (1) ⇒ (2).
= m−2 1 − n−m
2 2 X is compact ⇒ X has Balzano - Weierstrass property.
1
< m−2 → 0 as m → ∞ and n → ∞. ⇐⇒ X is sequentially compact (using Theorem 4.2.7).
2
(2) ⇒ (3).
Since, X is complete and {xn } is Cauchy sequence in X.
X is sequentially compact ⇒ X is complete and totally bounded (using
lim xn = x, (where x ∈ X),
n→∞ Theorem 4.2.12).
since x ∈ X and G is an open cover of X, x ∈ G for some G ∈ G. (3) ⇒ (1).
⇒ x ∈ S(x; ϵ) ⊆ G. X is complete and totally bounded ⇒ X is compact (using Theorem
Consider, Sk for any k and z ∈ Sk . 4.2.13).
By construction, Sk can not be covered finitely by G.
Corollary 4.2.15. Let (X, d) be a complete metric space. Then a
But,
closed subset Y of X is compact if and only if Y is totally bounded.
d(x, z) ≤ d(x, xk ) + d(xk , z)
1 Proof. Suppose X is complete and Y is a closed subset of X.
< ϵ + k , (allow k → ∞)
2
⇒ Y is also complete.
< ϵ, for large value of k or z ∈ S(x; ϵ)
54 55
By hypothesis, Y is totally bounded. Using Theorem 4.2.14, Y is if for every ϵ > 0, ∃ δ > 0 such that d(x, x′ ) < δ
compact.
⇒ |f (x) − f (x′ )| < ϵ, ∀ x, x′ ∈ X, f ∈ F.
Suppose Y is compact ⇒ Y is totally bounded (again using Theorem
4.2.14). Also, a family of functions F in C(X) is said to be uniformly bounded,
if ∃ M > 0 such that |f (x)| ≤ M , ∀ x ∈ X and f ∈ F.
Corollary 4.2.16. Let (X, d) be a complete metric space and let Y ⊆
X. Then Y is compact if and only if Y is totally bounded. Note 4.2.18. (1) f ∈ F and F is equi-continuous family
⇒ f is uniformly continuous on X.
Proof. Suppose, Y is compact ⇒ Y is complete and totally bounded.
n (2) If K is the set of all uniformly bounded functions, then K is
S
⇒ ϵ > 0, ∃ A = {a1 , a2 , . . . , an } ⊆ X such that Y ⊆ Sϵ (ak )
k=1 a bounded set in C(X). The uniformly boundedness is nothing but
n
S
⇒Y ⊆Y ⊆ Sϵ (ak ) boundedness in C(X).
k=1
∴ Y is also totally bounded.
Theorem 4.2.19 (Ascoli-Arzela Theorem). Let (X, d) be a com-
Suppose, Y is totally bounded ⇒ Y is also totally bounded (using (2)
pact metric space. Let C(X) be the complete metric space of all contin-
of Observation 4.2.11) since Y is a closed subset of a complete space
uous, bounded, real or complex valued functions on X with the metric
X.
is given by supremum norm. Then a closed subset K of C(X) is com-
⇒ Y is also complete.
pact ⇐⇒ K is equi-continuous and uniformly bounded.
Since, Y is complete and totally bounded, Y is compact (using Theo-
rem 4.2.14). Proof. Let us apply the Theorem 4.2.14 to K. Then K is compact if
and only if K is complete and totally bounded. But C(X) is complete
Definition 4.2.17. Let (X, d) be a metric space. Let C(X) be the
and K is a closed subset of C(X). So, K is complete. As a result, we
complete metric space of all continuous, bounded, real or complex val-
obtain K is compact if and only if K is totally bounded.
ued functions on X with the metric is given by the supremum norm
Suppose K is compact. First we shall show that K is equi-
||f || = sup |f (x)| and d(f, g) = ||f − g||.
x∈X continuous.
56 57
⇒ K has an 3ϵ -net, A = {f1 , f2 , . . . , fN } Suppose K is equi-continuous and uniformly bounded.
⇒ for every f ∈ K, |f (x) − fi (x)| < 3ϵ , i = 1, 2, . . . , N . We know that C(X) is complete metric space and K is a closed subset
Since, fi ’s continuous and K is compact, fi is also uniformly continu- of C(X), we shall show that K is totally bounded.
ous on K. By definition of uniformly continuity, given ϵ > 0, ∃ δi > 0 K is equi-continuous ⇒ given ϵ > 0, ∃ δ > 0 such that |f (x)−f (x′ 0)| <
ϵ
such that d(x, x′ ) < δi 4 whenever d(x, x′ ) < δ, ∀ x, x′ ∈ X, f ∈ K.
≤ |f (x) − fi (x)| + |fi (x) − fi (x′ )| + |fi (x′ ) − f (x′ )| −M = y0 < y1 < y2 < · · · < yk = M
58 59
Claim: E is an ϵ-net. 2. A. E. Taylor - Introduction to Functional Analysis, Willey, New
York, 1958.
|f (x) − ei (x)| = |f (x) − f (xi ) + f (xi ) − yij + yij − ei (xi ) + ei (xi ) − ei (x)|
≤ |f (x) − f (xi )| + |f (xi ) − yij | + |e(xi ) − yij | + |ei (x) − ei (xi )| −−−−−−−−−−−−−−−−−−−−−
ϵ ϵ ϵ ϵ
< + + + =ϵ
4 4 4 4
n
S
∴f ∈K⇒f ∈ S(ei , ϵ).
i=1
n
S
∴K⊆ S(ei , ϵ) or K is totally bounded. Hence K is compact.
i=1
4.3 Exercises
4.4 References
60 61
Block - II
Chapter 5
5.1 Introduction
62 63
operations in the form of linear expressions namely, vector addition space with respect to the co-ordinatewise linear operations defined in
and scalar multiplication. Starting from any vectors x, y, ∈ X, the Example 2.
vector addition enables to find the third vector z = x+y ∈ X. Starting (5) Let P be set of all polynomials, with real coefficients, defined on
from a scalar α and a vector x, the scalar multiplication enables to the closed unit interval [0, 1]. We specifically include all non-zero con-
find the vectors αx ∈ X. stant polynomials (which have degree 0) and the polynomial which is
identically zero (this has no degree at all). If the linear operations are
Definition 5.2.1. The Linear space X, written as (X, +,˙, F ) satsfies
taken to be the usual addition of two polynomials and the multiplica-
the following properties.
tion of a polynomial by a real number, then P is a real linear space.
(1) (X, +) is an abelian group.
(6) For a given positive integer n, let Pn be subset of P consisting of
(2) (a) α ∈ F, x, y ∈ X =⇒ α(x+y) = αx +αy.
the polynomial which is identically zero and all polynomials of degree
(b) α,β ∈ F , x ∈ X =⇒ (α + β)x = αx + βx.
less than n. Pn is a real linear space with respect to the linear opera-
(c) α, β ∈ F, x ∈ X =⇒ α(βx) = (αβ)x.
tions defined in P.
(d) x ∈ X =⇒ 1.x = x.
(7) A linear space may consist solely of the vector 0, with scalar mul-
The elements of X are called vectors and X is called Linear space.
tiplication defined by α.0 = 0 for all α. We refer to this as the zero
Example 5.2.2. (1) The set R of real numbers, with ordinary addi- space, and we always denote it by {0}.
tion and multiplication taken as the linear operations, is a real linear
Definition 5.2.3. A nonempty subset Y ⊂ X is said to be a subspace
space.
of X = (X, +, ., F ), if (Y, +, ., F ) is also a linear space.
(2) The set Rn of all n-tuples of real numbers is a real linear space
⇐⇒ (i) y1 , y2 ∈ Y =⇒ y1 + y2 ∈ Y
under the following co-ordinatewise linear operations: if
(ii) α ∈ F, y ∈ Y =⇒ αy ∈ Y
x = (x1 , x2 , ..., xn ) and y = (y1 , y2 , ..., yn ),
⇐⇒ α, β ∈ F, y1 , y2 ∈ Y =⇒ αy1 + βy2 ∈ Y . As a consequence, if α
then x + y = (x1 + y1 , x2 + y2 , ..., xn + yn ) and αx = (αx1 , αx2 , ..., αxn ).
= 1, β = −1, y1 = y2 , then αy1 + βy2 = y1 − y2 = 0 ∈ Y.
(3) The set C n of all complex numbers is a complex linear space under
ordinary addition and multiplication. Note 5.2.4. The necessary condition for Y to be a subspace of X is
(4) The set C n of all n-tuples of complex numbers is a complex linear 0∈Y.
64 65
Definition 5.2.5. Let X be a linear space and Y ⊂ X be a subspace is called a linear combination of x1 , x2 , . . . , xn .
of X. Then the quotient space of X with respect to Y is given by
Definition 5.2.8. If S is an arbitrary nonempty subset of linear space
X/Y = {x + Y : x ∈ X}.
X, then the set of all linear combinations of vectors in S is clearly
Equality of two numbers : x1 + Y = x2 + Y ⇐⇒ x1 − x2 ∈ Y .
a subspace of X; we denote this subspace by [S], and we call it the
Vector addition: (x1 + Y ) + (x2 + Y ) = (x1 + x2 ) + Y .
subspace spanned by S.
Scalar multiplication: α(x + Y ) = αx + Y .
Zero element: (x + Y ) + (z + Y ) = (x + Y ) Definition 5.2.9. If M is a subspace of linear space X, then a nonempty
66 67
that X is the sum of the subspaces M and N . This means that each N and they are equal; it therefore follows from M ∩ N = 0 that both
vector in X expressible as the sum of a vector in M and a vector in sides are 0, that x1 = x2 and y1 = y2 and that the decomposition of z
N . The case in which even more is true namely, that each vector z is unique.
in X is expressible uniquely in the form z = x + y, with x ∈ M and
Theorem 5.2.13. If S is a linearly independent set of vectors in a
y ∈ N will be of particular importance for us. In this case we say that
linear space X, then there exists a basis B for X such that S ⊂ B.
X is the direct sum of the subspaces M and N , and we symbolize this
statement by writing X = M ⊕ N . Proof. Consider the class P of all linearly independent subsets of X
which contain X. P is clearly a partially ordered set with respect to
Theorem 5.2.12. Let a linear space X be the sum of two subspaces
set inclusion. It suffices to show that P contains a maximal set B,
M and N , so that X = M + N . Then X = M ⊕ N ⇔ M ∩ N = {0}.
for such a maximal set will automatically be a basis for X such that
Proof. We begin by assuming that X = M ⊕ N and we deduce a S ⊆ B. By Zorn’s lemma ( i.e., If P is a partially ordered set in which
contradiction from the further assumption that there is a nonzero every chain has an upper bound, then P has a maximal element),
vector z in M ∩ N . It suffices to observe that z is expressible in it suffices to show that every chain in P has an upper bound in P.
two different ways as the sum of a vector x in M and a vector y in N , But this is evident from the fact that the union of all the sets in any
for z = z + 0 (here x = z and y = 0) and z = 0 + z (here x = 0 and chain of linearly independent sets which contain S is itself a linearly
y = z). This contradicts the uniqueness required by the assumption independent set which contains S.
that X = M ⊕ N .
We now assume that M ∩ N = {0} and we show that it follows 5.3 Linear operators
from X = M + N can be strengthened to X = M ⊕ N . Since X =
M + N , each z in X can be written in the form z = x + y with x ∈ M In this section, we define the linear operators and a theorem related
have two such decompositions of z, so that z = x1 + y1 = x2 + y2 , then Definition 5.3.1. Let N and N ′ be two normed linear spaces over the
x1 − x2 = y2 − y1 . The left side of this is in M , the right side is in same field F (= R or C). A mapping T : N −→ N ′ is said to be a
68 69
linear operator from N into N ′ , if Given: T (xn ) → 0 as xn → 0
T (αx + βy) = αT (x) + βT (y), ∀ x, y ∈ N and α, β ∈ F . =⇒ T (xn − x) → 0 as xn → x → 0 or xn → x
=⇒ (T (xn ) − T (x)) → 0 as xn → x
Definition 5.3.2. A linear operator T : N −→ N ′ is said to be con-
=⇒ T (xn ) → T (x) as xn → x, ∀ x ∈ N .
tinuous on N , if T (xn ) −→ T (x) as xn −→ x, ∀ x ∈ N .
=⇒ T is a continuous linear operator.
Definition 5.3.3. A linear operator T : N −→ N ′ is said to be (2) =⇒ (3) or not(3) =⇒ not(2)
bounded linear operator, if Suppose (3) is false =⇒ for every n > 0, ∃ xn ∈ N such that
||T (x)|| ≤ α||x|| for some α > 0, and ∀ x ∈ N . ||T (xn )|| > n||x|| .
xn xn 1
Theorem 5.3.4. Let T : N −→ N ′ be a linear operator from a normed Set yn = n||xn || . Then ||yn || = || n||x n ||
|| = n||xn || ||xn || = n1 .
1
linear space N into a normed linear space N ′ . Then the following are As n → ∞, yn → 0 (∵||yn || =⇒ n → 0 as n → ∞).
equivalent:
xn
But, T (yn ) = T n||x n ||
T (xn )
= n||x n ||
=⇒ ||T (yn )|| = ||Tn||x
(xn )||
n ||
> 1, ∀ n.
(4) If S = {x ∈ N : ||x|| ≤ 1} is the closed unit sphere in N , then (3) =⇒ T is a bounded linear operator
70 71
=⇒ for s ∈ S = {x ∈ N : ||x|| ≤ 1}, ||T (s)|| ≤ α||s|| ≤ α (2) Each of the following conditions determines a subset of the real
=⇒ ||T (s)|| ≤ α, ∀ s ∈ S linear space C[−1, 1] of all bounded continuous real functions y = f (x)
=⇒ d(T (s)) ≤ 2α < ∞, (d(A) = diameter of A = sup d(a, b)) defined on [−1, 1]:
a,b∈A
=⇒ T (s) is bounded in N ′ =⇒ (4). (a) f is differentiable;
(4) =⇒ (3) (b) f is polynomial of degree 3;
′
(4) =⇒ If S = {x ∈ N : ||x|| ≤ 1}, then T (s) is bounded in N (c) f is an even function, in the sense that f (−x) = f (x) for all x;
=⇒ d(T (s)) = α < ∞ (d) f is an odd function, in the sense that f (−x) = −f (x) for all x;
=⇒ ||T (s)|| ≤ α, for s ∈ S (e) f (0) = 0;
x
=⇒ 0 ̸= x ∈ N , s = ||x|| , ||T (s)|| ≤ α (f) f (0) = 1;
=⇒ ∃ α > 0 such that ||T x
|| ≤ α (g) f (x) ≥ 0 for all x.
||x||
=⇒ α > 0 such that ||T (x)|| ≤ α||x||, ∀ x ∈ N Which of these subsets are subspaces of C[−1, 1]?
5.5 References
(1) Each of the following conditions determines a subset of the real 2. A. E. Taylor - Introduction to Functional Analysis, Willey, New
3
linear space R of all triples x = (x1 , x2 , x3 ) of real numbers: York, 1958.
(a) x1 is an integer;
−−−−−−−−−−−−−−−−−−−−−
(b) x1 = 0 or x2 = 0;
(c) x1 + 2x2 = 0;
(d) x1 + 2x2 = 1.
Which of these subsets are subspaces of R3 ?
72 73
(3) ||λx|| = |λ|||x||
(4) ||x|| = 0 implies x = 0.
In this unit, we introduce the concept of Normed spaces and bounded (2) C(S) (the set of real or complex valued continuous functions on S)
operators. is a normed linear space with the norm
||f || = sup{|f (x)| : x ∈ S}.
6.2 Normed Spaces Proposition 6.2.5. Every normed linear space X is metric space,
relative to the natural metric d defined by
We begin by restating the definition of Normed spaces. d(x, y) = ||y − x||, ∀ x, y ∈ X.
Definition 6.2.1. Let X be a real or complex linear space and F be Furthermore, for any x, y, z ∈ X and for all λ ∈ F , we have
a field of real or complex numbers. A norm on X is a real function ||x|| = d(0, x) as well as
|| · || : X −→ R defined on X such that for any x, y ∈ X and for all (a) d(x + z, y + z) = d(x, y), (i.e. translation invariance).
(1) ||x|| ≥ 0
Proof. It can be easily verified that the axioms for a metric hold good.
(2) ||x + y|| ≤ ||x|| + ||y||
For example: d(x, z) ≤ d(x, y)+d(y, z) follows immediately by writing
74 75
z − x = (y − x) + (z − y) integer k, there is a positive integer nk such that
so that ||z − x|| ≤ ||y − x|| + ||z − y||. 1
||xn − xm || < , ∀ n, m ≥ nk .
Proof of (a): d(x + z, y + z) = ||(y + z) − (x + z)|| = ||y − x|| = d(x, y). 2k
Proof of (b): d(λx, λy) = ||λy − λx|| = ||λ(y − x)|| = |λ|||y − x|| = Choose nk+1 > nk . Let y1 = xn1 and yk+1 = xnk+1 − xnk , k ≥ 1.
∞
|λ|d(x, y).
P
Then ||yk || < ∞. Therefore, there exists y ∈ X such that
k=1
Theorem 6.2.6. A normed linear space X is complete if and only if m
X
y = lim yk = lim xnm .
every absolutely convergent series in X is convergent. m→∞ m→∞
k=1
Conversely, Suppose every absolutely convergent series in X is Since N ′ is a linear space over F , B(N, N ′ ) is also a linear space over
convergent. Let {xn } be a Cauchy sequence in X. For each positive F.
76 77
Theorem 6.3.1. The linear space B(N, N ′ ) over F is a normed linear (3)
space with respect to ||T1 + T2 || = sup{||(T1 + T2 )(x)|| : ||x|| ≤ 1}
||T || = sup{||T (x)|| : ||x|| ≤ 1}.
= sup{||T1 (x) + T2 (x)|| : ||x|| ≤ 1}
Proof. To show that || · || is a norm on B(N, N ′ ): ≤ sup{||T1 (x)|| + ||T2 (x)|| : ||x|| ≤ 1}
= ||T1 || + ||T2 ||
||T || = 0 ⇐⇒ sup{||T (x)|| : ||x|| ≤ 1} = 0
⇐⇒ if s ∈ S = sup{x ∈ N : ||x|| ≤ 1}, then ||T (s)|| = 0 B(N, N ′ ) is a normed linear space.
x x
⇐⇒ if x ∈ N, then ∈ S and hence T = 0 (x ̸= 0)
||x|| ||x||
1 6.4 Exercises
⇐⇒ T (x) = 0, ∀ x ∈ N (|| · ||N ′ is norm on N ′ )
||x||
⇐⇒ T = 0. (1) Prove that R* is a normed linear space with following norms
n
P
(a) ||x||1 = |xi |.
(2) i=1
(b) ||x||∞ = max |xi |.
1≤i≤n
||λT || = sup{||(λT )(x)|| : ||x|| ≤ 1} (2) Prove that C[a, b] is a normed linear space with ||f || = sup |f (x)|.
x∈[a,b]
= sup{|λ|||T (x)|| : ||x|| ≤ 1}
−−−−−−−−−−−−−−−−−−−−−
78 79
Definition 7.2.1. Let P be a nonempty set. A partial order relation in
P is a relation symbolized by ≤ and it satisfies the following properties:
(1) Reflexivity: x ≤ x, ∀ x ∈ P .
(2) Anti-symmetry: x ≤ y and y ≤ x =⇒ x = y, ∀ x, y ∈ P .
Chapter 7 (3) Transitivity: x ≤ y, y ≤ z then x ≤ z, ∀ x, y, z ∈ P .
Then (P, ≤) is called a partially ordered set.
Unit 7: Hahn-Banach Extension Definition 7.2.2. A partially ordered set (P, ≤) is said to be totally
ordered set. If x, y ∈ P, then either x ≤ y or y ≤ x.
Theorem
Definition 7.2.3. Let (P, ≤) be a totally ordered set. Then y ∈ P is
called a maximal element, if x ≥ y ⇒ x = y.
7.1 Introduction Definition 7.2.4. Let (P, ≤) be a totally ordered set and A is a
nonempty subset of P , then y ∈ P is called an upper bound for A,
In this unit, we shall prove the Hahn-Banach Extension Theorem.
if x ≤ y, ∀ x ∈ A.
This is one of the most fundamental theorems in functional analysis
Definition 7.2.5. If P is a partially ordered set in which every chain
and is due to Hahn and Banach. It yields the existence of nontrivial
has an upper bound, then P has a maximal element
continuous linear functionals on a normed linear space, a basic result
OR
necessary for the development of a large portion of functional analysis.
If P is a totally ordered set in which every subset has an upper bound,
There are several forms of this famous theorem. We shall first prove
then P has a maximal element.
the extended form of the Hahn-Banach Theorem.
Definition 7.2.6. Let X be a real linear space. By a sublinear func-
80 81
Definition 7.2.7. Let X be a complex linear space. By a seminorm Since f ∈ F, F ̸= ∅.
on X, we mean p : X → R satisfying the following properties: Let us define a partial order “≤” in F as follows:
(1) p(x + y) ≤ p(x) + p(y), ∀ x, y ∈ X. g1 ≤ g2 ⇐⇒ D(g1 ) ⊆ D(g2 ) and g1 (x) = g2 (x) on x ∈ D(g1 ).
(2) p(αx) = |α|p(x), ∀ x ∈ X, α ∈ C. (a) “≤” is reflexive:
g ≤ g because D(g) = D(g), g(x) = g(x) (obvious).
Note 7.2.8. (1) If we restrict C to only R, then every seminorm is
(b) “≤” is antisymmetric:
also a sublinear functional.
g1 ≤ g2 and g2 ≤ g1 =⇒ D(g1 ) ⊆ D(g2 ) ⊆ D(g1 ) or D(g1 ) = D(g2 ),
(2) Every norm, by definition, is also a seminorm. Any result proved
g1 (x) = g2 (x) on D(g1 ) = D(g2 ) =⇒ g1 = g2 .
on seminorm will hold for norms.
(c) “≤” is transitive:
(3) For any seminorm, p(x) ≥ 0.
Suppose g1 ≤ g2 and g2 ≤ g3 =⇒ D(g1 ) ⊆ D(g2 ) ⊆ D(g3 ) or
Theorem 7.2.9 (Hahn-Banach Theorem for any real linear D(g1 ) ⊆ D(g3 ), g1 (x) = g2 (x), ∀ x ∈ D(g1 ), or g2 (x) = g3 (x), ∀ x ∈
space). Let X be a real linear space. Let S be a subspace of X. Let D(g3 ) =⇒ g1 (x) = g3 (x), ∀ x ∈ D(g1 )
p be a sublinear functional on X. Let f : S → R be a real linear ∴ g1 ≤ g3 .
functional on S such that f (x) ≤ p(x), ∀ x ∈ S. Then there exists a Hence (F, ≤) is a partially ordered set.
real linear functional F : X → R such that: Let G be a subcollection of F such that (G, ≤) is a totally ordered set.
(1) F is an extension of f or F (x) = f (x), ∀ x ∈ S and Claim G has an upper bound or there exists an extension h of f such
domain of F ⊇ domain of f . that g ≤ h, ∀ g ∈ G.
(2) F (x) ≤ p(x), ∀ x ∈ X. Set U =
S
D(g). U is a subspace of X.
g∈G
Because, x, y ∈ U and α, β ∈ R.
Proof. Let F denote the collection of all extensions of f
=⇒ x ∈ D(g1 ) and y ∈ D(g2 ) for some g1 , g2 ∈ G.
i.e. g ∈ F =⇒
=⇒ since G is totally ordered, either g1 ≤ g2 or g2 ≤ g1 .
(i) g(x) = f (x), ∀ x ∈ S and D(f ) ⊆ D(g), (where D(f ) = domain of
Without loss of generality, g1 ≤ g2 .
f ).
=⇒ x ∈ D(g1 ) ⊆ D(g2 ) and y ∈ D(g2 )
(ii) g(x) ≤ p(x), ∀ x ∈ D(g).
82 83
=⇒ x, y ∈ D(g2 ), D(g2 ) is a linear subspace of X. Hence h is an upper bound of F or G.
=⇒ αx + βy ∈ D(g2 ) ⊆ U . We have proved that in the partially ordered set (F, ≤), every totally
=⇒ αx + βy ∈ U . ordered subset has an upper bound. By Zorn’s Lemma, F has a
∴ U is a subspace of X. maximal element. Let F : X → R be that maximal element i.e.,
Suppose x ∈ U , then x ∈ D(g) for some g ∈ G may be more than one (1) F is an extension of f .
domain. (2) F (x) ≤ p(x), ∀ x ∈ D(F ).
If x ∈ D(g1 ) and x ∈ D(g2 ) for some g1 , g2 ∈ G or x ∈ D(g1 ) ∩ D(g2 ), The proof will be complete if, we show that D(F ) = X.
then, again we can use the fact that G is totally ordered and assume We shall show that, if D(F ) ̸= X, then F is not a maximal element.
without loss of generality, g1 ≤ g2 or D(g1 ) ⊆ D(g2 ) which will lead to If D(F ) ̸= X, then set D = Span{D(F )∪{x0 }}, where x0 ∈ X\ D(F ).
x ∈ D(g1 ) ∩ D(g2 ) = D(g1 ). For every x ∈ D, we may write x = y + αx0 , where y ∈ D(F ), α∈ R.
Hence, every x ∈ U falls or lies exactly in a precise domain of a ∀ y, z ∈ D(F ),
members g1 ∈ G.
F (z) − F (y) = F (z − y)
As a result, define h : U → R by h(x) = g(x) whenever x ∈ D(g) ⊆ U
≤ p(z − y) = p(z + x0 − y − x0 )
(g is fixed according to the above discussion).
≤ p(z + x0 ) + p(−y − x0 )
h is well defined. h is linear because
or − F (y) − p(−y − x0 ) ≤ p(z + x0 ) − F (z).
h(αx + βy) = g(αx + βy)
= f (x). and
84 85
Then a < b and choose c such that a < c < b. As a result g(x) ≤ p(x).
−F (y) − p(−y − x0 ) ≤ c ≤ p(z + x0 ) − F (z). This shows that there is one more maximal element superseding F or
Now construct a real linear function g : D −→ R by g(x) = F (y) + αc, F is not a maximal element. This is impossible.
(x = y + αx0 ), since F is well defined and linear, g is also well defined ∴ D(F ) = X. This completes the proof.
and linear. Theorem 7.2.10 (Hahn-Banach Theorem for any complex lin-
To show that g(x) ≤ p(x), ∀ x ∈ D. ear space). Let X be a complex linear space. Let S be a subspace of
Case 1: α = 0. X. Let p : X −→ R be a seminorm on X. Let f : S −→ C be a linear
g(x) = g(y) = F (y) ≤ p(y) = p(x) functional on S such that ∥f (x)∥ ≤ p(x), ∀ x ∈ S. Then there exists
∴ g(x) ≤ p(x). a complex linear functional F : X −→ C such that:
Case 2: α > 0. (1) F (x) = f (x), ∀ x ∈ S.
c ≤ p(z + x0 ) − F (z), ∀ z ∈ D(F ). (2) |F (x)| ≤ p(x), ∀ x ∈ X.
y
Take z = α,
Proof. First, we observe that f (x) = g(x) + ih(x), ∀ x ∈ S,
y y
c ≤ p( + x0 ) − F ( ), (α > 0)
α α where g : S −→ R and h : S −→ R are two real linear functionals.
y y
αc ≤ αp( + x0 ) − αF ( ) Next, we want to express h(x) in terms of g(x):
α α
= p(y + αx0 ) − F (y) if (x) = ig(x) − h(x)
= p(x) − F (y) if (x) = f (ix),
86 87
g(x) ≤ |f (x)| ≤ p(x), ∀ x ∈ S, (∵ z = Re(z) + iIm(z), Re(z) ≤ |z|).
So, by restricting C to R, X is also a real linear space. S is a subspace
F ((α1 + iα2 )x) = F (α1 x + iα2 x)
of X. p : S −→ R is a sublinear functional. g : S −→ R is a linear
= G(α1 x + iα2 x) − iG(i(α1 x + iα2 x))
functional such that g(x) ≤ p(x).
= G((α1 + iα2 )x) − iG(iα1 x − α2 x)
By applying Hahn-Banach Theorem for any real linear space, ∃ a real
linear functional, G : X −→ R such that = G(α1 x) + G(iα2 x) − iG(iα1 x − α2 x)
= F (x) + F (y). Finally, to show that |F (x)| ≤ p(x), ∀ x ∈ X, (∵ z = x+iy = |z| = eiθ )
Note that, F (x) = |F (x)|eiθ
or |F (x)| = F (x)e−iθ
= F (xe−iθ ), ∵ F is linear
88 89
=⇒ Im F (xe−iθ ) = 0.
(1)
= p(x) + p(y)
∴ |F (x) ≤ p(x), ∀ x ∈ X.
This completes the proof. or p(x + y) ≤ p(x) + p(y), ∀ x, y ∈ X.
90 91
x |F (x)|
∥F ∥ = sup F : x ∈ X = sup :x∈X
∥x∥ ∥x∥
|F (x)|
≥ sup :x∈S ∵ S is a subspace of X.
∥x∥
= sup
|f (x)|
: x ∈ S , ∵ F (x) = f (x), ∀ x ∈ X
Chapter 8
∥x∥
= ∥f ∥
92 93
If A ⊆ C(X, R) and A is also a lattice, then we call A as sublattice of constant function, then A = C(X, R).
C(X, R).
Proof. Step 1: To show that A is a closed sublattice of C(X, R). First,
we observe that if f ∈ C(X, R), |f | ∈ C(X, R).
8.2 Stone-Weierstrass Theorem Here, |f |(x) = |f (x)|, ∀ x ∈ X.
Next, we note that
In order to state the result we need the following definition.
(f ∨ g)(x) = max{f (x), g(x)}
Definition 8.2.1. C(X, R) is said to separate points of X, if x, y ∈ X, (f + g) + |f − g|
=
x ̸= y, then ∃ f ∈ C(X, R) such that f (x) ̸= f (y). 2
Let ϵ > 0, we know that |t| is a continuous function on [a, b]. By Consider the open set Gf = {z : f (z) < h(z) + ϵ}.
Weierstrass theorem (i.e., let f : [a, b] → R be a continuous function Keep x fixed and vary y in X. Then the corresponding f is denoted
and let ϵ > 0 be given. Then ∃ a polynomial p(x) ∈ R[x], such that by fy and the corresponding Gf is denoted by Gfy or simply Gy .
|f (x) − p(x)| < ϵ), there exist a polynomial p(t) such that Then, Gy = {z ∈ X : fy (z) < h(z) + ϵ} is an open set, ∀ y ∈ X.
S
||t| − p(t)| < ϵ. =⇒ X = Gy. But X is compact, ∃, y1 , y2 , . . . , yn ∈ X such that
y
n
Take [a, b] = [−||f ||, ||f || ], since A is a subalgebra of C(X, R),
S
X= G yi .
i=1
p(f ) ∈ A ⊆ C(X, R). Correspondingly, in L, ∃ f1 , f2 , . . . , fn such that
Then, ||f (x)| − p(f (x))| < ϵ, ∀ x ∈ X. fi (z) < h(z) + ϵ, ∀ i = 1, 2, . . ..
=⇒ || |f (x)| − p(f (x)) || < ϵ. Let gx = f1 ∧f2 ∧· · · wedgefn . Infact gx (x) = h(x) and gx (z) < f (z)+ϵ,
Hence |f (x)| ∈ A = A. ∀ z ∈ X.
Hence A is a closed sublattice of C(X, R). Next, consider for every x, the set Hx = {z ∈ X : gx (z) > f (z) − ϵ}
(hence, f (z) − ϵ < gx (z) < f (z) + ϵ, ∀ z ∈ X). Hx is an open set in
fying the property: if x, y are two distinct points of X and if a, b are ∴X=
[
Hx .
two real numbers, then ∃ f ∈ L such that f (x) = a and f (y) = b, x
m
then L = C[a, b].
S
X is compact, hence ∃ x1 , x2 , . . . , xm ∈ X such that X = Hx i .
i=1
Since L is a closed sublattice of C(X, R), L ⊆ C(X, R). Correspondingly, ∃ g1 , g2 , . . . , gm ∈ L such that
To show that C(X, R) ⊆ L or h ∈ C(X, R) =⇒ h ∈ L. f (z) − ϵ < gi (z) < f (z) + ϵ, ∀ i = 1, 2, . . . , m.
96 97
Let gϵ = g1 ∨ g2 ∨ · · · ∨ gm . Then gϵ ∈ L such that function and contains conjugates of each of its functions. Then A
h(z) − ϵ < gϵ (z) < h(z) + ϵ or |h(z) − gϵ (z)| < ϵ. equal to C(X, C).
Theorem 8.2.4 (Stone-Weierstrass Theorem for complex case). Corollary 8.2.5. Weierstrass theorem on C[a, b] is a corollary to
Let X be a compact Hausdorff space and A be closed subalgebra of Stone-Weierstrass theorem in real case.
C(X, C), which separates the points of X, contains non zero constant
98 99
Proof. X = [a, b] =⇒ C[a, b] = C(X, R) X = [a, b] × [c, d] in the Euclidean plane R2 , then f can be uniformly
P [a, b] = {p(x) ∈ C[a, b] : p(x) = a0 + a1 x + · · · + an xn , a1 , a2 , . . . , an ∈ approximated on X by polynomials in x and y with real coefficients.
R}.
The nonzero constant function 1 ∈ P [a, b]. The function f (x) = x ∈ (2) Let X be the closed unit disc in the complex plane, then show
P [a, b], obviously separates points of X = [a, b]. that any function f in C(X, C) can be uniformly approximated on X
A = P [a, b] is a closed subalgebra of C([a, b], R) containing nonzero by polynomials in z and z with complex coefficients.
constant function 1 and f (x) = x separates points of [a, b].
∴ C[a, b] = A = P [a, b]. This completes the proof.
(3) Let X and Y be compact Hausdorff spaces and f a function in
Note 8.2.6. P (A, B) is dense in C[a, b] C(X ×Y, C). Show that f can be uniformly approximated by functions
n
=⇒ given ϵ > 0 and f ∈ C[a, b], ∃ pϵ ∈ P [A, B] such that fi gi where the fi ’s are in C(X, C) and the gi ’s are in
P
of the form
i=1
|f − pϵ | < ϵ. C(Y, C).
This is called Classical Weierstrass Approximation theorem.
Proof. X = Sr (x0 ) = {x ∈ Rn : ||x − xo|| ≤ r} 1. G. F. Simmons - Introduction to Topology and Modern Analysis,
C(X, R), A = P (X, R), closures of the space of all polynomials in n- Tata McGraw-Hill, New Delhi, 2004.
variables with real coefficients.
2. A. E. Taylor - Introduction to Functional Analysis, Willey, New
A = C(X, R).
York, 1958.
100 101
Block-III
Chapter 9
9.1 Introduction
102 103
with the norm is ∥x∥ = |x|. Proof.
(2) X = Rn or Cn = F n is a Banach space with the norm is n
X
!1/p n
X
!1/q
P q
p P ut α = |xj | and β = |yj | .
x = (x1 , x2 , . . . , xn ) =⇒ ∥x∥ = |x1 |2 + |x2 |2 + · · · + |xn |2 . j=1 j=1
!1/p
xj yj
x
X Put x′j = α and yj′ = β,
(3) X = lp = x = (x1 , x2 , . . . , xn ) : |xj |p < ∞, xj ∈ F = R or C
j=1
n
!1/p n
!1/p n
!1/p
X 1 X 1 X 1
is a Banach space with the norm is |x′j |P = |xj |P = |xj | P
= α = 1.
j=1
αp j=1 α j=1
α
∞
!1/p
X
p
∥ x ∥p = |xj | . n
X
!1/q
j=1 Similarly |yj′ |q = 1.
1 1 a p bq j=1
Lemma 9.2.3. If p > 1, q > 1, p+q = 1 and a, b > 0 then ab < p+q.
Consider,
Proof. n n ′ p
|xj | |yj′ |q
X X
Z a
ap |x′j yj′ | ≤ + , using Lemma 9.2.3
A1 = xp−1 dx = . j=1 j=1
p q
0 p
n n
! !
1 X ′ P 1 X ′ q
= |x | + |y |
p j=1 j q j=1 j
y = xp−1 ⇐⇒ x = y 1/p−1 = y q−1 .
1 1
= + = 1.
b
p q
bq
Z
A2 = y q−1 dy = .
0 q
n
ap b q X xj yj
∴ ab < A1 + A2 = + . ⇒ ≤1
p q α β
j=1
n
1 X
⇒ |xj yj | ≤ 1
Theorem 9.2.4 (Holder’s Inequality). Let (x1 , x2 , . . . , xn ) and (y1 , y2 , . . . , yn ) αβ j=1
be n-tuples. Then n
X n
X
!1/p n
X
!1/q
p q
n n
!1/p n
!1/q ⇒ |xj yj | ≤ αβ = |xj | |yj |
X X X
P q j=1 j=1 j=1
|xj yj | ≤ |xj | |yj |
j=1 j=1 j=1 This completes the proof.
1 1
where p > 1, q > 1 and p + q = 1.
104 105
Problem 9.2.5. Let C ∗ (X) = {f : X → F : F = R or C}, where f C ∗ (X) is complete.
is bounded continuous function on X, with Let {fn } be a Cauchy sequence in C ∗ (X).
vector addition: (f + g)(x) = f (x) + g(x), ∀x ∈ X. ∥ fm − fn ∥ = sup {|(fm − fn )(x)|}
x∈X
scalar multiplication: (αf )(x) = αf (x), ∀ x ∈ X
= sup {|fm (x) − fn (x)|}.
x∈X
and norm
=⇒ {fn (x)} is a Cauchy sequence in F and F is complete.
∥ f ∥= sup |f |.
x∈X Hence {fn (x)} → f (x) as n → ∞, ∀ x ∈ X.
Since, F is linear space on itself, C ∗ (X) is also a linear space on F .
∥ fm − fn ∥≥ |(fm (x) − fn (x))|
Show that C ∗ (X) is Banach space.
or |(fm (x) − fn (x))| ≤∥ fm − fn ∥→ 0 as m, n → ∞.
106 107
9.3 Exercises
108 109
T : X −→ Y is continuous linear map from X onto Y , then the image Next, let us note that
∞
of every open sphere centered on the origin of X contains an open [
X = Sn .
sphere centered on the origin of Y . n=1
∞
!
[
=⇒ T (x) = T Sn
Proof. Let Sr = {x ∈ X : ∥x − 0∥x < r} be an open sphere in X, n=1
∞
centered on origin of X. [
= T (Sn ).
Let Sr′ = {y ∈ Y : ∥y − 0′ ∥Y < r} be an open sphere in Y , centered n=1
x
Next, consider z ∈ T (Sn0 ) − y0 .
Put r = x1
=⇒ z = Tx − y0 , where x ∈ Sn0
Sr = {rx1 ∈ X : ∥x1 ∥ < 1} =⇒ z = Tx − Tx0 , ∴ y0 ∈ T (Sn0 ) =⇒ ∃ x0 ∈ Sn0 such that
= rS1 T (x0 ) = y0 .
=⇒ z = T (x − x0 ).
Therefore, Sr = rS1 , ∀ r > 0.
∥x − x0 ∥ ≤ ∥x∥ + ∥x0 ∥ < n0 + n0 = 2n0 , ∴ x − x0 ∈ S2n0 .
From this, we note that it is enough to prove the Lemma for S1 .
′ ∴ z ∈ T (S2n0 )
i.e. to prove that, T (S1 ) ⊇ Sr1 , for some r1 > 0.
and hence,
T (Sn0 ) − y0 ⊆ T (S2n0 ).
We shall observe that origin of Y is an interior point of T (Sn0 ) − y0 .
The map y −→ y − y0 is a homeomorphism from Y onto Y . Since
110 111
ϵ
y0 is an interior point of T (Sn0 ), 0′ = y0 − y0 is an interior point of 2.
But, we know that T (S2n0 ) = T (2n0 S1 ) = 2n0 T (S1 ), ∵ Sr = rS1 . Continuing this way we arrive at two sequences: {xn } is a sequence in
So that, T (S2n0 ) = 2n0 T (S1 ), X such that
1
we can have one more homeomorphism y −→ 2n0 y from Y onto Y . ∥x∥ < , n = 1, 2, . . . .
2n−1
Hence origin 0′ is an interior point of T (S1 ). and yn is a sequence in Y such that ∥y − y1 − y2 − · · · − yn ∥ < ϵ
2n .
′
By definition, ∃ ϵ > 0 such that 0 ∈ Sϵ′ ⊂ T (S1 ) using this, we shall
=⇒ y = lim (y1 + y2 + · · · + yn )
show that n→∞
Sϵ′ ′
⊆ T (S3 ) ⇐⇒ S ϵ ⊆ T (S1 ), = lim (T (x1 ) + T (x2 ) + · · · + T (xn ))
3
n→∞
We shall work out inductively two sequences {yn } in Y and {xn } = lim (T (Sn )),
n→∞
in X as follows: where Sn = x1 + x2 + · · · + xn .
Let y ∈ Sϵ′ , then y ∈ T (S1 ) or y is a limit point of T (S1 )) (if y ∈
T (S1 ) ⊆ T (S3 ) there is nothing to prove). n
X
=⇒ ϵ
∃ x1 ∈ S1 such that y1 = T (x1 ) and ∥y − y1 ∥ < 2ϵ , ∥Sn ∥ = xi
2,
i=1
∵ x1 ∈ S1 ⇐⇒ ∥x1 ∥ < 1. n
X
ϵ ≤ ∥xi ∥
=⇒ ∥(y − y1 ) − 0′ ∥ < 2 or y − y1 ∈ S ′ϵ using 0′ ∈ Sϵ′ ⊆ T (S1 ), we get i=1
2
=⇒ y − y1 ∈ T (S 21 ) or y − y1 is a limit point of T (S 21 ).
ϵ 1
=⇒ 22 , ∃ x2 ∈ S 21 or ∥x∥ < 2 such that y2 = T (x2 ) and ∥y −y1 −y2 ∥ <
112 113
Banach spaces and if T : X −→ Y is a continuous linear map from X
1 1 onto Y , then T is an open mapping.
∥Sn ∥ < 1 + + · · · + n−1
2 2
1 − 21n Proof. y ∈ T (G) =⇒ y = T (x) for some x ∈ G.
=
1 − 12
< 2. G is open in X =⇒ ∃ r > 0 such that x ∈ S(x, r) ⊆ G.
But we know that S(x, r) = x + Sr .
To show that {Sn } is a Cauchy sequence in X:
By the Lemma 10.2.3, ∃ r1 > 0 such that Sr′ 1 ⊆ T (Sr ).
n
X
∥Sn − Sm ∥ = xi S ′ (y, r1 ) = y + Sr′ 1 ⊆ y + T (Sr ) = T (x + Sr ) = T (S(x, r)).
i=m+1
n ∴ y ∈ S ′ (y, r1 ) ⊆ T (S(x, r)) ⊆ T (G)
X
≤ ∥xi ∥ ∴ T (G) is open.
i=m+1
1 1 1
< m
+ m+1 + · · · + n−1 Theorem 10.2.5. If T : X −→ Y is a one to one, continuous linear
2 2 2
1
1 1 − 2n−m transformation (or bounded linear operator) from X onto Y , then T
= m
2 1 − 12
is a homeomorphism from a Banach space X onto Banach space Y .
1 1
= m−1 1 − n−m
2 2
Proof. It is given that T is a continuous linear transformation, one to
1
= m−1 −→ 0 as n, m −→ ∞
2 one and onto from a Banach space X to a Banach space Y .
∴ {Sn } is a Cauchy sequence in X and X is complete. By Open Mapping Theorem, G ⊆ X is open =⇒ T (G) ⊆ Y is
Let x = lim Sn . also open.
n→∞
Since T is continuous, T (x) = limn→∞ T (Sn ) = y.
=⇒ T (G) ⊆ Y is open =⇒ T −1 (T (G)) = G ⊆ X is open.
∥x∥ = lim ∥Sn ∥ ≤ 2 < 3.
n→∞
=⇒ T −1 is continuous.
=⇒ x ∈ S3
=⇒ y = T (x) ∈ T (S3 ) or S ′ (ϵ) ⊆ T (S3 ) or S ′ (ϵ/3) ⊆ T (S1 ). T −1 (y) = T −1 (T (x)) = x.
114 115
10.3 Exercise:
T is one to one and onto =⇒ T −1 is also one to one and onto. 10.4 References
∴ T is homeomorphism.
1. G. F. Simmons - Introduction to Topology and Modern Analysis,
Problem 10.2.6. If T : X −→ Y is a one to one continuous linear Tata McGraw-Hill, New Delhi, 2004.
transformation from a Banach space X onto Banach space Y , then
2. A. E. Taylor - Introduction to Functional Analysis, Willey, New
T −1 : Y −→ X is also continuous linear transformation.
York, 1958.
= T −1 (T (αx1 + βx2 ))
= αx1 + βx2
= αT −1 (y1 ) + βT −1 (y2 ).
116 117
Note 11.2.3. Every continuous linear operator is also a closed linear
operator. But the converse is not true.
Counter example
Chapter 11 Let
Unit 11: Closed Graph Theorem Y = {y : [0, 1] −→ R : ∥y∥ = sup |y(t)|, y(t) is continuous.}
0≤t≤1
dx ′
T : X −→ Y , by Tx (t) = dt = x (t).
T is linear:
11.1 Introducation
d
T (αx1 (t) + βx2 (t)) = [αx1 (t) + βx2 (t)]
dt
In this unit, we introduce the concept of Closed graph theorem. dx1 dx2
= α +β
dt dt
= αTx1 + βTx2 .
11.2 Closed Graph Theorem
T is closed:
Suppose xn −→ x in X and Txn −→ y in Y .
First, we define some definition and then go to the theorem.
i.e. xn (t) −→ x(t) in X and x′n (t) −→ y(t) ∈ Y.
Definition 11.2.1. Let T : X −→ Y be a linear transformation from
=⇒ ∥xn − x∥ −→ 0 and ∥x′n − y∥ −→ 0 as n −→ ∞.
a normed linear space X into a normed linear space Y . Then T is
=⇒ xn (t) −→ x(t) uniformly and x′n (t) −→ y(t) uniformly on [0, 1].
said to be a closed linear operator if
By the theorem on uniform convergence and differentiation,
xn −→ x in X and Txn −→ y in Y =⇒ y = Tx .
x′n −→ x(t) uniformly and hence y = x′ (t).
Definition 11.2.2. Let T : X −→ Y be a linear transformation from ∴ y = Tx or T is a closed linear operator.
a normed linear space X into a normed linear space Y . Then T is But T is not continuous.
said to be continuous if xn −→ x in X and Txn −→ Tx . ( T is continuous ⇐⇒ ∃ α > 0 such that ∥Tx ∥ ≤ α∥x∥ ⇐⇒ x′ (t) is
118 119
uniformly continuous.) T is defined by
Consider, xn (t) = tn .
Γ(T ) = {(x, T (x)) : x ∈ X} ⊆ X × Y.
120 121
Theorem 11.2.8 (Closed Graph Theorem). Let X and Y be two X × Y is also a normed linear space under the norm
Banach spaces and let T : X −→ Y is a linear transformation from
∥(x, y)∥ = ∥x∥ + ∥y∥.
X into Y . Then T is continuous ⇐⇒ Γ(T ) is closed.
To show that X × Y is also complete.
Proof. Suppose T is continuous. Let (x, y) be a limit point of Γ(T ). Let {(xn , yn )} be a Cauchy sequence in X × Y .
Then by definition of limit point, ∃ a sequence {(xn , T (xn ))} in Γ(T ) =⇒ ∥(xm , ym ) − (xn − yn )∥ −→ 0 as m, n −→ ∞
such that lim (xn , T (xn )) = (x, y). =⇒ ∥(xm − xn , ym − yn )∥ −→ 0 as m, n −→ ∞
n→∞
=⇒ ∥(xn , T (xn )) − (x, y)∥ −→ 0 as n −→ ∞. =⇒ ∥xm − xn ∥X −→ 0 and ∥ym − yn ∥Y −→ 0 as m, n −→ ∞.
=⇒ ∥(xn − x, T (xn ) − y)∥ −→ 0 as n −→ ∞. =⇒ {xn } and {yn } are Cauchy sequences in X and Y respectively.
=⇒ ∥xn − x∥ + ∥T (xn ) − y∥ −→ 0 as n −→ ∞. Also, X and Y are complete.
=⇒ ∥xn − x∥ −→ 0 and ∥T (xn ) − y∥ −→ 0 as n −→ ∞. =⇒ ∃ x ∈ X, y ∈ Y such that lim xn −→ x and lim yn −→ y.
n→∞ n→∞
(T is continuous ⇐⇒ xn −→ x =⇒ T (xn ) −→ T (x)) =⇒ lim (xn , yn ) = (x, y) ∈ X × Y .
n→∞
=⇒ As xn −→ x, T (xn ) −→ y and T (xn ) −→ T (x). ∴ X × Y is a Banach space.
=⇒ y = T (x). Next, X × Y is complete and Γ(T ) is a closed subspace of X × Y .
∴ (x, y) = (x, T (x)) ∈ Γ(T ). =⇒ Γ(T ) is also complete.
Γ(T ) has all its limit points or Γ(T ) = Γ(T ). =⇒ Γ(T ) is also a Banach space.
∴ Γ(T ) is closed. Define a map, ϕ : Γ(T ) −→ X by ϕ(x, T (x)) = x.
Suppose Γ(T ) is closed subspace of X × Y . It is given that X and Y To show that ϕ is linear.
are Banach spaces.
ϕ(α(x1 , T (x1 )) + β(x2 , T (x2 ))) = ϕ((αx1 + βx2 , αT (x1 ) + βT (x2 )))
To show that X × Y is also a Banach space:
= ϕ((αx1 + βx2 , T (αx1 + βx2 )))
X × Y = {(x, y) : x ∈ X, y ∈ Y } is a vector space on common field F
under the following operations: ∵ T is linear.
122 123
∴ ϕ is linear. Solution: Let X and Y be Banach spaces.
To show that ϕ is continuous: Given the Closed graph theorem is true. i.e if T : X −→ Y is a linear
∥ϕ((x, T (x)))∥ = ∥x∥ ≤ ∥x∥ + ∥T (x)∥ = ∥(x, T (x))∥. transformation, then T is continuous ⇐⇒ Γ(T ) is closed.
∴ ϕ is continuous. By definition, ϕ is one to one and onto. In the theorem, ϕ : X −→ Y is a one to one, continuous linear mapping
∴ ϕ is one to one continuous linear mapping from the Banach space from X onto Y .
Γ(T ) onto the Banach space X. To prove that ϕ−1 is one to one, linear, continuous and onto:
=⇒ By Theorem 10.2.5., ϕ−1 is also a one to one continuous linear ϕ−1 is continuous ⇐⇒ Γ(ϕ−1 ) is closed.
mapping from X onto Γ(T ). Γ(ϕ−1 ) = {(y, ϕ−1 y) : y ∈ Y }.
=⇒ ∥ϕ−1 ∥ = α > 0, is well defined. Let (y, x) be a limit point of T (ϕ−1 )
Consider, =⇒ ∃ (yn , ϕ−1 (yn )) in Γ(ϕ−1 ) such that lim (yn , ϕ−1 (yn )) = (y, x)
n→∞
=⇒ ∥(yn , ϕ−1 (yn )) − (y, x)∥ −→ 0 as n −→ ∞.
∥T (x)∥ ≤ ∥x∥ + ∥T (x)∥
=⇒ ∥(yn − y, ϕ−1 (yn ) − x)∥ −→ 0 as n −→ ∞
= ∥(x, T (x))∥
=⇒ ∥yn − y∥ −→ 0 and ∥ϕ−1 (yn ) − x∥ −→ 0 as n −→ ∞.
= ∥ϕ−1 (x)∥, ∵ ϕ((x, T (x))) = x =⇒ (x, T (x)) = ϕ−1 (x).
Since ϕ is onto, yn = ϕ(xn ).
≤ ∥ϕ−1 ∥∥x∥
=⇒ ∥ϕ(xn ) − y∥ −→ 0 and ∥ϕ−1 (ϕ(xn )) − x∥ −→ 0 as n −→ ∞.
= α∥x∥. =⇒ ∥xn − x∥ −→ 0 and ∥ϕ(xn ) − y∥ −→ 0 as n −→ ∞.
Since ϕ is continuous, ϕ is closed. ϕ(xn ) −→ ϕ(x) and y = ϕ(x) or
∴ ∥T (x)∥ ≤ α∥x∥ or T is continuous. This completes the proof.
x = ϕ−1 y
(ϕ : X −→ Y is closed if xn −→ x and ϕ(xn ) −→ y, then y = ϕ(x)).
Problem 11.2.9. To prove closed graph theorem we have used one of ∴ ϕ−1 (yn ) −→ x = ϕ−1 (y) as yn −→ y.
the consequence of open mapping theorem, namely, if ϕ : X −→ Y is ∴ (y, x) = (y, ϕ−1 y) ∈ Γ(ϕ−1 )
a one to one continuous linear mapping from a Banach space X onto ∴ Γ(ϕ−1 ) is closed.
a Banach space Y , then ϕ−1 : Y −→ X is also one to one continuous By Closed graph theorem, ϕ−1 is continuous.
linear mapping from Y onto X.
124 125
ϕ is linear, one to one and onto =⇒ ϕ−1 is linear, one to one and onto.
This completes the proof.
11.3 Exercises
Chapter 12
If T : X −→ Y is a closed linear operator from a normed space X into
a normed linear space Y , then show that Γ(T ) is a closed subspace of
Unit 12: Banach - Steinhaus
X ×Y.
Principle of Uniform Boundedness
11.4 References
126 127
Proof. For each positive number n, define Next, let y ∈ B and ∥y∥ ≤ 1,
z
Fn = {x ∈ B : ∥Ti (x)∥ ≤ n, ∀ i} =⇒ ∥Ti (y)∥ = Ti , where z = r0 y =⇒ ∥z∥ < r0 .
r0
1
\
= Ti−1 (S(0, n)). = ∥Ti (z)∥
i r0
1
= ∥Ti ((z + x0 ) − x0 )∥
r0
1
≤ [∥Ti (z + x0 )∥ + ∥Ti (x0 )∥].
Because, x ∈ Fn ⇐⇒ Ti (x) ∈ S(0, n) in N. r0
By Baire’s Category Theorem, B is of second category. So, ∃ n0 such Problem 12.2.2. Let B be a Banach space and N be a normed linear
that space. Let {Tn } be a sequence in B(B, N ) such that lim Tn (x) = T (x)
n→∞
Int(Fn0 ) = Int(Fn0 ) ̸= ∅. exists for all x ∈ B. Show that T ∈ B(B, N ).
128 129
3. {Tn } is a sequence of continuous linear transformation from B to Given: f ∈ N ∗ =⇒ f : N −→ F is a continuous linear transforma-
N. tion.
=⇒ f : N −→ F is a bounded linear transformation.
4. lim Tn (x) = T (x) exists for all x ∈ B.
n→∞
=⇒ ∥f ∥ < ∞.
From (4), ∥Tn (x) − T (x)∥ −→ 0asn −→ ∞, ∀x ∈ B. Consider,
=⇒ ∥(Tn − T )(x)∥ −→ 0 as n −→ ∞, ∀ x ∈ B. diam(f (X)) = sup |f (x) − f ′ (x)|
x,x′ ∈X
=⇒ {(Tn − T )(x)} is a bounded sequence of N for all x ∈ B.
We can apply principle of uniform boundedness on {(Tn − T )(x)} to
≤ 2 sup |f (x)| < ∞
x∈X
get {∥Tn − T ∥} is a bounded set of numbers.
provided |f (x)| < ∞, ∀ x ∈ X.
=⇒ ∥Tn − T ∥ ≤ α, ∀ x ∈ B.
|f (x)| ≤ ∥f ∥∥x∥ ≤ ∥f ∥diam(X) < ∞
=⇒ ∥(Tn − T )(x)∥ ≤ α, ∀ x ∈ B
=⇒ diam(f (x)) < ∞.
=⇒ ∥Tn (x) − T (x)∥ ≤ α∥x∥, ∵ |a − b| ≥ |a| − |b|.
∴ f (X) is bounded set of numbers ∀ f ∈ N ∗ .
=⇒ ∥T (x) − Tn (x)∥ ≤ α∥x∥
=⇒ ∥T (x)∥ − ∥Tn (x)∥ ≤ α∥x∥ Suppose f (X) is a bounded set of numbers ∀ f ∈ N ∗ .
Since Tn is continuous ∃ αn > 0 such that ∥Tn (x)∥ ≤ αn ∥x∥. Let us write X = {xi } (i need not run over countably infinite) for our
130 131
property that {Ti (x)} is bounded set of N1 for every x ∈ B. Then To show that T ∗ is continuous:
{∥Ti ∥} is bounded set of numbers i.e. {Ti } is bounded in B(B, N ).
∥T ∗ (f )∥ = sup{|[T ∗ (f )](x)| : ∥x∥ ≤ 1}
Set B = N ∗ , N1 = N ∗∗ and {Ti } = {Fxi }.
= sup{|f (T (x))| : ∥x∥ ≤ 1}
Then by principle of uniform boundedness, {Fxi } is bounded set
= ∥f T ∥
of numbers. But by the constructions, ∥Fxi ∥ = ∥xi ∥.
≤ ∥f ∥∥T ∥, ∵ ∥T1 T2 ∥ ≤ ∥T1 ∥∥T2 ∥.
Hence {∥xi ∥} is a bounded set of numbers or X is bounded in N .
This completes the proof. Since T ∈ B(N ), take α = ∥T ∥ < ∞.
∴ ∥T ∗ (f )∥ ≤ α∥f ∥
Theorem 12.2.4. Let N be a normed linear space. Let T ∈ B(N )
It remains to show that ∥T ∗ ∥ = ∥T ∥:
i.e. T : N −→ N is a continuous linear transformation on N. Define
T ∗ : N ∗ −→ N ∗ by (T ∗ (f ))(x) = f (T (x)). Then T ∗ is also a continu- ∥T ∗ ∥ = sup{∥T ∗ (f )∥ : ∥f ∥ < 1}
ous linear transformation from N to F or T ∗ (f ) ∈ N ∗ , ∀ f ∈ N ∗ . = ∥T ∥, (∥T ∗ (f )∥ ≤ α∥f ∥).
Further, T −→ T ∗ is an isometric isomorphism from B(N ) into B(N ∗ ).
132 133
12.3 Properties of T ∗ 12.4 Exercises
Proposition 12.3.1. (1) T ∗ is also a continuous linear transforma- 1. Let B be a Banach space and N a normed linear space. If {Tn }
tion from N into F is a sequence of in B(B, N ) such that T (x) = lim Tn (x) exists
n→∞
for each x in B, then prove that T is a continuous linear trans-
(2) ∥T ∗ ∥ = ∥T ∥ (Isometry)
formation.
(3) (T1 T2 )∗ = T2∗ T1∗ (reversing the product property).
2. Let T be an operator on a normed linear space N . If N is consid-
∗
(4) I = I (Preserving the identity). ered to be part of N ∗∗ by means of the natural embedding, then
show that T ∗∗ is an extension of T . Further show that if N is
Proof. Proof of (1) and (2) follows from the Theorem 12.2.4.
reflexive, then T ∗∗ = T .
Proof of (3):
= [(T2∗ T1∗ )(f )](x), ∀ x ∈ N, f ∈ N ∗ . 2. A. E. Taylor - Introduction to Functional Analysis, Willey, New
Proof of (4):
−−−−−−−−−−−−−−−−−−−−−
∗
[I (f )](x) = f (I(x))
= f (x)
= [I(f )](x)
∴ I ∗ = I.
134 135
Block-IV
Chapter 13
13.1 Introduction
136 137
Proof. We verify the following three properties of the norm: Hence H is a normed linear space.
p
1. ∥x∥ = (x, x) ≥ 0 (∵ (x, x) ≥ 0 ) for all x ∈ H, and
∥x∥ = 0 ⇐⇒ ∥x2 ∥ = 0
13.2 Hilbert space
⇐⇒ (x, x) = 0
Definition 13.2.1. A Banach space (or a complete normed linear
⇐⇒ x = 0, (∵ (x, x) = 0 ⇐⇒ x = 0, ∀ x ∈ H).
space) is said to be Hilbert space, if the norm is defined by the inner
2. product. In other words, a complete inner product space defining a
p Banach space is called a Hilbert space.
∥λx∥ = (λx, λx)
q
= λλ(x, x) Example 13.2.2.
∞
p ( )
= |λ2 |∥x∥2 X
2
l2 = x = (x1 , x2 , . . . , xn , . . .) : |xn | < ∞, xn ∈ C, n = 1, 2, . . . .
= |λ|∥x∥, ∀ λ ∈ F, x ∈ H. n=1
=⇒ ∥x + y∥ ≤ ∥x∥ + ∥y∥ ∀ x, y ∈ H.
138 139
The series on the right hand side is absolutely convergent by Holder’s 2. A. E. Taylor - Introduction to Functional Analysis, Willey, New
inequality York, 1958.
∞
X ∞
X
xn yn = |xn yn |
n=1 n=1
−−−−−−−−−−−−−−−−−−−−−
∞
!1/2 ∞
!1/2
X X
≤ |xn |2 |yn |2
n=1 n=1
< ∞.
13.3 Exercises:
( 1/2 )
∞
2
P
1. Show that l2 = x = (x1 , x2 , . . . , xn , . . .) : |xn | < ∞, xi ∈ F .
n=1
is a Hilbert space.
Rb
2. Show that L2 (I) = {f : I −→ R or C : a |f |2 dx < ∞} is a
Hilbert space.
13.4 References
140 141
Theorem 14.1.2. In a Hilbert space H, the inner product is contin-
uous. i.e. if xn −→ x, yn −→ y as n −→ ∞, then (xn , yn ) −→ (x, y)
as n −→ ∞.
Chapter 14
Proof.
Unit 14: Properties of Hilbert |(xn , yn ) − (x, y)| = |(xn , yn ) − (xn , y) + (xn , y) + (x, y)|
= |(xn , yn − y) + (xn − x, yn )|
Spaces
≤ |(xn , yn − y)| + |(xn − x, yn )|
= ∥xn ∥ × 0 + 0 × ∥y∥ = 0 as n −→ ∞.
In this section, we shall give the some properties of the Hilbert spaces.
This completes the proof.
Theorem 14.1.1. In a Hilbert space, the parallelogram law holds:
Note 14.1.3. (1) lim (xn , y) = lim xn , y = (x, y).
n→∞ n→∞
∥x + y∥2 + ∥x − y∥2 = 2(∥x∥2 + ∥y∥2 ).
(2) lim (x, yn ) = x, lim yn = (x, y).
n→∞ n→∞
Proof. Consider,
(3) lim (xn , yn ) = (x, y).
n→∞
2 2 2
∥x + y∥ + ∥x − y∥ = (x + y, x + y) + (x − y, x − y), using ∥x∥ = (x, x)
Theorem 14.1.4. In a Hilbert space H,
= (x, x) + (x, y) + (y, x) + (y, y)
+(x, x) − (x, y) − (y, x) + (y, y) 4(x, y) = ∥x + y∥2 − ∥x − y∥2 + i∥x + iy∥2 − i∥x − iy∥2 ,
√
= 2(x, x) + 2(y, y) where i = −1.
= 2∥x∥2 + 2∥y∥2 = 2 ∥x∥2 + ∥y∥2 .
142 143
Proof. Consider, 2. A. E. Taylor - Introduction to Functional Analysis, Willey, New
York, 1958.
∥x + y∥2 − ∥x − y∥2 + i∥x + iy∥2 − i∥x − iy∥2
= 4(x, y).
14.2 Exercise
14.3 References
144 145
Theorem 15.2.2. In a Hilbert space H, if x ⊥ y, then
.
Chapter 15
Proof.
∥x + y∥2 = (x + y, x + y)
Unit 15: Orthogonal projection
= (x, x) + (x, y) + (y, x) + (y, y)
and nearly orthogonal elements = (x, x) + (x, y) − (y, x) + (y, y)
= ∥x∥2 + ∥y∥2 .
∥x − y∥2 = (x − y, x − y)
In this unit, we introduce the concept of orthogonal projection and
= (x, x) − (x, y) − (y, x) + (y, y)
give the properties of orthogonality.
= (x, x) − (x, y) + (y, x) + (y, y)
= ∥x∥2 + ∥y∥2 .
15.2 Orthogonal projection
146 147
Note 15.2.4. (1) If S = H, then S ⊥ = {0}. To show that S ⊥ is a closed subspace of H :
(2) If S = {0}, then S ⊥ = H. Let x be a limit point of S ⊥ . Then ∃ {xn } in S ⊥ such that lim xn = x.
n→∞
Consider, for all y ∈ S,
Theorem 15.2.5. Let H be a Hilbert space and S ⊆ H. Then S ⊥
satisfies the following properties: (x, y) = lim xn , y = lim (xn , y) = 0.
n→∞ n→∞
(1) S ∩ S ⊥ = {0}.
∴ x ∈ S ⊥.
(2) S ⊥ is a closed subspace of H.
Hence S ⊥ has all of its limit points or S ⊥ is a closed subspace of H.
(3) S1 ⊆ S2 =⇒ S2⊥ ⊆ S1⊥ .
(4) S ⊆ S ⊥⊥ .
(3) Suppose, S1 ⊆ S2 . To show that S2⊥ ⊆ S1⊥ .
Proof. (1) Let x ∈ S2⊥ . Then (x, y) = 0, ∀ y ∈ S2 .
x ∈ S ∩ S ⊥ ⇐⇒ x ∈ S and x ∈ S ⊥ =⇒ (x, y) = 0, ∀ y ∈ S2 ⊇ S1 .
=⇒ (x, y) = 0, ∀ y ∈ S1 .
⇐⇒ (x, x) = 0
∴ x ∈ S1⊥ .
⇐⇒ x = 0
Hence S2⊥ ⊆ S1⊥ .
⇐⇒ x ∈ {0}
148 149
Theorem 15.2.6. Let M be a closed linear subspace of a Hilbert space =⇒ {xn } is a Cauchy sequence in M and {yn } is Cauchy sequence in
H. Then H = M ⊕ M ⊥ . N.
Every closed subspace of a complete space is complete =⇒ M, N are
Proof. Step - 1: First, let us prove that if M, N are two closed sub-
complete.
spaces of a Hilbert space H and M ⊥ N , then M + N is also a closed
=⇒ lim xn = x, lim yn = y, where x + y ∈ M + N .
n→∞ n→∞
suspace of H.
∴ M + N is also a closed subspace of H.
Let α, β ∈ F, z1 , z2 ∈ M + N
Now, it is given M is a closed subspace of H and M ⊥ is always
=⇒ αz1 + βz2 = α(x1 + y1 ) + β(x2 + y2 ) = (αx1 + βx2 ) + (αy1 + βy2 )
a closed subspace of H. Hence M + M ⊥ is a closed subspace of H.
(where z1 = x1 + y1 and z2 = x2 + y2 , x1 , x2 ∈ M , y1 , y2 ∈ N ).
Further, M ∩ M ⊥ = {0}. As a result, M + M ⊥ = M ⊕ M ⊥ .
Since αx1 + βx2 ∈ M , being a subspace of H and αy1 + βy2 ∈ N ,
being a subspace of H. We must have αz1 + βz2 ∈ M + N .
Hence M + N is a subspace of H. Step 2: To show that H = M + M ⊥ .
=⇒ ∃ a sequence {zn } in M + N such that lim zn = z. of H. Then there exist 0 ̸= z0 ∈ H such that z0 ⊥ M + M ⊥ . Then
n→∞
Since M ⊥ N , M ∩N = {0}. As a result zn ∈ M +N =⇒ zn = xn +yn z0 ⊥ M and z0 ⊥ M ⊥⊥ orz0 ∈ M ⊥ ∩ M ⊥⊥ = {0} or 0 ̸= z0 = 0, which
Pythagoras theorem,
Definition 15.2.7. Let H be a Hilbert space. A non empty set {ei }
150 151
thonormal set. Definition 15.2.10. Let H be a Hilbert space. Let P denote the class
of all orthonormal sets in H. Under set inclusion ⊆, (P, ⊆) is a
Theorem 15.2.9 (Bessel’s inequality). If {ei } is an orthonormal
partially ordered set. By Zorn’s lemma, there exists a maximal element
set in Hilbert space H and if x ∈ H, then
n for the chain of orthonormal sets in H. This set or the maximal
X
2 2
|(x, ei )| ≤ ∥x∥ . orthonormal set is called complete orthonormal set.
i=1
Proof. First, let us observe that, Theorem 15.2.11. Let {ei } be an orthonormal set in Hilbert space
n
X
2 H. Then the following are equivalent:
x− (x, ei )ei ≥0
(1) {ei } is complete;
i=1
n n
!
X X (2) x ⊥ {ei } =⇒ x = 0;
=⇒ x− (x, ei )ei , x − (x, ei )ei ≥0 P
i=1 i=1 (3) x ∈ H =⇒ x = (x, ei )ei (called Fourier expansion of x);
n n
! !
(4) x ∈ H =⇒ ∥x∥2 = |(x, ei )|2 (called Parsval’s identity).
X X P
=⇒ (x, x) − x, (x, ei )ei − (x, ei )ei , x
i=1 i=1
n n
!
X X Proof. (1) =⇒ (2):
+ (x, ei )ei , (x, ei )ei ≥0
i=1 i=1 Suppose if possible, x ⊥ {ei } and x ̸= 0.
n n n
x
Put e = ∥x∥ . Then {ei , e} is also an orthonormal set containing {ei }.
X X X
=⇒ ∥x∥2 − (x, ei )(x, ei ) − (x, ei )(ei , x) + (x, ei )(x, ei ) ≥ 0
i=1 i=1 i=1
n n n =⇒ {ei } is not complete, a contradiction.
X X X
2 2 2 2
=⇒ ∥x∥ − |(x, ei )| + |(x, ei )| + |(x, ei )| ≥ 0. ∴ x = 0.
i=1 i=1 i=1
n
X
∴ |(x, ei )|2 ≤ ∥x∥2 , if {ei } is finite.
(2) =⇒ (3)
i=1
n
X
|(x, ei )|2 ≤ ∥x∥2 .
i=1
152 153
Consider, 15.3 Exercises
X X
x− (x, ei )ei , ej = (x, ei ) − (x, ei )(ei , ej )
1. Show that the parallelogram law is not true in l1n (n > 1).
= (x, ei ) − (x, ei )
2. If S is a nonempty subset of a Hilbert space, then show that
= 0, ∀ ej .
S ⊥ = S ⊥⊥⊥ .
P
=⇒ x − (x, ei )ei ⊥ {ei }.
P P
∴ x − (x, ei )ei = 0 or x = (x, ei )ei . 15.4 References
(4) =⇒ (1)
Suppose (1) is false,
=⇒ ∃ e ∈ H such that e ⊥ {ei } and ∥e∥ = 1.
=⇒ ∥e∥2 = |(e, ei )|2 .
P
=⇒ 1 = 0, a contradiction.
This completes the proof.
154 155
Since M is closed, d must be strictly greater than zero. For oth-
erwise d = 0 implies x ∈ M from the property closed sets, contracting
the choice of x.
156 157
Theorem 16.3.2 (Riesz representation theorem). Let H be a f (x)
f (x) − βf (y0 ) = 0 and this is accomplished by putting β = f (y0 ) . Our
∗
Hilbert space and let f be an arbitrary functional in H . Then there conclusion that f (x) = (x, y) is true for every x in H now follows at
exists a unique vector y ∈ H such that f (x) = (x, y) for every x in H. once from
Proof. It is easy to see that if such a y exists, then it is necessarily f (x) = f (m + βy0 )
′ ′
unique. For if we also have f (x) = (x, y ) for all x, then (x, y ) = (x, y) = f (m) + βf (y0 )
′
and (x, y − y) = 0 for all x; and since 0 is the only vector orthogonal = (m, y) + β(y0 , y)
to every vector, this implies that y ′ − y = 0 or y ′ = y.
= (m + βy, y)
We now turn to the problem of showing that y does exist. If = (x, y).
f = 0, then it clearly suffices to choose y = 0. We may therefore
This completes the proof.
assume that f ̸= 0. The null space M of f is thus a proper closed
linear subspace of H, then there exists a non-zero vector y0 which
is orthogonal to M . We show that if α is a suitable chosen scalar, 16.4 Exercises
then the vector α may be, f (x) = (x, y) is true for every x in M ; for
1. Let H be a Hilbert space. Then show that H ∗ is also a Hilbert
f (x) = 0 for such an x and since x is orthogonal to y0 , we also have
space with respect to the inner product defined by (fx , fy ) =
(x, y) = 0. This allows us to focus our attention on choosing α in such
(y, x). In just the same way, the fact that H ∗ is a Hilbert space
a way that f (x) = (x, y) is true for x = y0 . The condition this imposes
implies that H ∗∗ is a Hilbert space whose inner product is given
on α is that
by (Ff , Fg ) = (f, g).
f (y0 ) = (y0 , αy0 ) = α∥y0 ∥2 .
f (y0 )
2. Let H be a Hilbert space. We have two natural mappings of H
We therefore choose α to be ∥y0 ∥2 and it follows that f (x) = (x, y)
into H ∗∗ , the second of which is onto: the Banach space natural
is true for every x in M and for x = y0 . It is easily seen that each
imbedding x −→ Fx , where Fx (f ) = f (x) and the product map-
x in M can be written in the form x = m + βy0 with m in M : all
ping x −→ fx −→ Ffx , where fx (y) = (y, x) and Ffx = (f, fx ).
that is necessary is to choose β in such a way that f (x − βy0 ) =
158 159
Such that these mappings are equal and conclude that H is re-
flexive. Show also that (Fx , Fy ) = (x, y).
16.5 References
−−−−−−−−−−−−−−−−−−−−−
160