Functional Analysis Notes
Functional Analysis Notes
2020
Contents
1
CONTENTS
2
0.1. VECTOR SPACE OR LINEAR SPACE
3
0.1. VECTOR SPACE OR LINEAR SPACE
1. x + y ∈ X
2. x + y = y + x
3. x + (y + z) = (x + y) + z
4. There exists an element 0 (called zero vector) such that x + 0 = 0 + x = x
5. For every x ∈ X, there exists an element −x, such that x + (−x) = 0 = (−x) + x
6. rx ∈ X
7. r(x + y) = rx + ry
8. (r + s)x = rx + sx
9. r(sx) = (rs)x
10. for all x ∈ X, 1.x = x.
Example 0.1.2. • R is a vector space over R with vector addition being the usual addition
of real numbers, and scalar multiplication being the usual multiplication of real numbers.
• Rn is a vector space over R, with addition and scalar multiplication defined as follows:
x1 y1 x1 y1 x1 + y1 x1
.
. .
. n
if . , . ∈ R , then . + . = .
. .
. .
. , if α ∈ R and ... ∈ Rn , then
.
x y xn yn xn + yn xn
n n
x1 αx1
.
α .. = .. .
.
xn αxn
• The function space C[a, b]. Let a, b ∈ R and a < b. Consider the vector space comprising
functions f : [a, b] → K that are continuous on [a, b], with addition and scalar multipli-
cation defined as follows. If f , g ∈ C[a, b], then f + g ∈ C[a, b] is the function given
by
( f + g)(x) = f (x) + g(x), x ∈ [a, b].
If α ∈ K and f ∈ C[a, b], then α f ∈ C[a, b] is the function given by (α f )(x) = α f (x),
x ∈ [a, b].
C[a, b] is referred to as a ‘function space’, since each vector in C[a, b] is a function (from
[a, b] to K).
• The sequence space s. Let s denote the set of all sequences x = {xn }∞ 1 of real or complex
numbers. Define the operations of addition and scalar multiplication pointwise: For
all x = (x1 , x2 , ...), y = (y1 , y2 , ...) ∈ s and all α ∈ K, define x + y = (x1 + y1 , ...) and
αx = (αx1 , ...). Then s is a linear space over K.
• The sequence space ` p , 1 ≤ p < ∞. Let ` p denote the set of all sequences x = {xn }∞1 of
real or complex numbers satisfying the condition ∑∞ |x | p < ∞, with addition and scalar
i=1 i
multiplication defined as follows:
1 + (yn )1 = (xn + yn )1 , (xn )1 , (yn )1 ∈ ` p ;
(xn )∞ ∞ ∞ ∞ ∞
4
0.1. VECTOR SPACE OR LINEAR SPACE
Proof. Let x = (x1 , x2 , ...), y = (y1 , y2 , ...) ∈ ` p . Now, we show that x + y ∈ ` p . Since for
each i ∈ N, we have
that is αx ∈ ` p .
• The sequence space `∞ . Let `∞ denote the vector space of all bounded sequences with
values in K, that is supi∈N |xi | < ∞ and with addition and scalar multiplication defined as
follows:
(xn )n∈N + (yn )n∈N = (xn + yn )n∈N , (xn )n∈N , (yn )n∈N ∈ `∞ ;
α(xn )n∈N = (αxn )n∈N , α ∈ K, (xn )n∈N ∈ `∞ .
• The sequence space c = c(N) Let c denote the set of all convergent sequences x = (xn )∞ 1
of real or complex numbers. That is, c is the set of all sequences x = (xn )∞ 1 such that
lim xn exists. Define the operations of addition and scalar multiplication pointwise as in
n→∞
previous example. Then c is a linear space over K.
• The sequence space c0 = c0 (N). Let c0 denote the set of all sequences x = (xn )∞
1 of real
or complex numbers which converge to zero. That is, c0 is the space of all sequences x =
(xn )∞ lim xn = 0. Define the operations of addition and scalar multiplication
1 such that n→∞
pointwise as in previous example. Then c0 is a linear space over field K.
5
0.2. LINEAR SUBSPACE
6
0.2. LINEAR SUBSPACE
and
f2 (x) = x4 + b1 x3 + c1 x2 + d1 x + e1 ∈ W
but f1 + f2 ∈ / W.
[ii] W = { f (x) : coeff of x2 = 2 or − 2} is not a subspace of V as f1 (x) = ax4 + bx3 +
2x2 + dx + e, f2 (x) = a1 x4 + b1 x3 − 2x2 + d1 x + e1 ∈ W but f1 + f2 ∈
/ W.
[iii] W = { f (x) : f (0) = 1} is not a subspace of V . Take f1 , f2 ∈ W such that f1 (0) =
1, f2 (0) = 1. Here ( f1 + f2 )(0) = 2 ∈/ W.
[iv] W = { f (x) : f (x) has negative coefficients} is not a subspace of V take α = −1
then α f ∈ / W.
[v] W = { f (x) : deg f (x) ≤ 3} is a subspace of V . Take f1 , f2 ∈ W such that degree of
f1 , f2 is ≤ 3. So α f1 + β f2 ∈ W for any α, β .
Definition 0.2.1. A linear combination of vectors x1 , x2 ..., xm of a vector space X is an expres-
sion of the form α1 x1 + α2 x2 + ... + αm xm , where the coefficients α1 , ..., αm are any scalars.
Definition 0.2.2. For any nonempty subset M ⊂ X, the set of all linear combinations of vectors
of M is called the span of M, written span M. [M] = {α1 x1 +α2 x2 +...+αm xm : αi ∈ F, ui ∈ M}
Obviously, this is a subspace Y of X, and we say that Y is spanned or generated by M.
Example 0.2.4. i) Let V3 = {(x1 , x2 , x3 ) : xi ∈ R, i = 1, 2, 3} be a vector space and M =
{(1, 0, 0), (1, 1, 1)} be any subset of V3 .
The span of set M is given by
But these not satisfied, so (1, 3, −5) do not belong to linear span generated by S.
7
0.3. LINEAR DEPENDENT/LINEAR INDEPENDENT
α1 x1 + α2 x2 + ... + αr xr = 0, (0.3.1)
a+b = 0
and
b = 0.
So a = 0 = b. Therefore M is linearly independent.
ii) Check whether set M = {x, |x|} is linearly independent or not in C (−1, 1) and C (1, 4).
Solution. Consider ax + b|x| = 0.
Case 1. If x ∈ (0, 1), then ax + bx = 0. As x 6= 0 so
a + b = 0. (0.3.2)
a − b = 0. (0.3.3)
8
0.3. LINEAR DEPENDENT/LINEAR INDEPENDENT
9
0.4. SOME INEQUALITIES
|a + b| p ≤ 2 p (|a| p + |b| p ).
Proof. Case 1: Take |a| > |b|. So we have |a + b| ≤ |a| + |b| < 2|a| and |a + b| p < 2 p |a| p ≤
2 p (|a| p + |b| p ).
Case 2. |a| ≤ |b|. So we have |a + b| ≤ |a| + |b| ≤ 2|b| and |a + b| p ≤ 2 p |b| p . Hence
|a + b| p ≤ 2 p (|a| p + |b| p ).
Here equality occurs if ai and bi are proportional, that is, ai = cbi , where c is constant.
Definition 0.4.1. Let p and q be positive real numbers. If 1 < p < ∞ and 1p + 1q = 1, or if p = 1
and q = ∞, or if p = ∞ and q = 1, then we say that p and q are conjugate exponents.
Lemma 0.4.3. (Young’s Inequality) Let p and q are conjugate exponents with 1 < p and α, β ≥
0. Then
αp βq
αβ ≤ + .
p q
Here, equality occurs iff β = α p−1 .
Proof. If p = 2 = q, then the inequality follows from the fact that (α − β )2 ≥ 0. Notice also,
that if α = 0 or β = 0, then the inequality follows trivially.
p q
If p 6= 2, then consider the function f : [0, ∞) → R given by f (α) = αp + βq − αβ , for fixed
β > 0.
1 q
Then, f 0 (α) = α p−1 − β = 0 when α p−1 = β , that is, when α = β p−1 = β p > 0.
q
We now apply the second derivative test to the critical point α = β p . f 00 (α) = (p −
q
1)α p−2 > 0, for all α ∈ (0, ∞). Thus, we have a global minimum at α = β p .
10
0.4. SOME INEQUALITIES
q p q p q
It is easily verified that 0 = f (β p ) ≤ f (α) = αp + βq − αβ ⇐⇒ αβ ≤ αp + βq , for each
α ∈ [0, ∞).
Or
The result follows directly if α = 0 or β = 0. Without the loss of generality, assume that
α > 0 and β > 0. Take p = mn + 1 and q = 1 + mn such that 1p + 1q = 1, where m and n are
positive integers. Assume α = x1/p and β = y1/q , we get
αp βq x y mx + ny
+ = n + m = .
p q m +1 1+ n m+n
We know that A.M.≥ G.M., we get
mx + ny 1
≥ (xm yn ) m+n = x1/p y1/q = αβ .
m+n
q
αp
Thus, αβ ≤ p + βq .
1
where p + 1q = 1, p > 1 and ∑∞ p ∞ q
i=1 |ai | < ∞, ∑i=1 |bi | < ∞.
1 1
|ai bi | ≤ |ai | p + |bi |q .
p q
Summing over i, we get
∞
1 ∞ 1 ∞ 1 1
∑ |aibi| ≤ p ∑ (|ai| ) + q ∑ (|bi|q) = p + q = 1.
p
i=1 i=1 i=1
and
bi
yn = 1 .
(∑∞ q q
i=1 |bi | )
Consider
|ai bi | |ai | p |bi |q
|xi yi | = 1 1 ≤ p
+ q
.
(∑∞ |a | p ) p (∑∞ |b |q ) q p ∑∞
i=1 |ai | q ∑∞
i=1 |bi |
i=1 i i=1 i
Summing over i, and using Young’s inequality we get
∞
|ai | |bi | 1 1
∑ |xiyi| = p
1
q
1 ≤ + = 1.
p q
i=1 (∑∞ ∞
i=1 |ai | ) (∑i=1 |bi | )
p q
1 1
Hence ∑∞ ∞ p p ∞ q q
i=1 |ai bi | ≤ (∑i=1 |ai | ) (∑i=1 |bi | ) .
11
0.4. SOME INEQUALITIES
Lemma 0.4.5. (Minkowski inequality) If p > 1 (fixed) and ai , bi are nonnegative real numbers
and ∑∞ p ∞ p
i=1 |ai | < ∞, ∑i=1 |bi | < ∞, then
∞ 1 ∞ 1 ∞ 1
( ∑ (|ai + bi |) p ) p ≤ ( ∑ |ai | p ) p + ( ∑ |bi | p ) p .
i=1 i=1 i=1
|cn | p = |an + bn | p = |an + bn ||an + bn | p−1 ≤ |an ||cn | p−1 + |bn ||cn | p−1 .
where q is such that 1/p + 1/q = 1 and q(p − 1) = p, 1 − 1/q = 1/p. Therefore, we have
m m m 1 m 1
( ∑ |cn | p )1/p = ( ∑ |an + bn | p )1/p ≤ ( ∑ |an | p ) p + ( ∑ |bn | p ) p .
n=1 n=1 n=1 n=1
Now note that this inequality holds for all m, and the series in the terms on the right-hand
side converge by assumption, so we can let m → ∞ to obtain the inequality we wanted.
12
0.5. METRIC SPACE
Here d(x, y) ≥ 0, d(x, y) = 0 iff x = y, d(x, y) = d(y, x). Also, for any case triangle
inequality holds.
(
|x − y|; x = cy; for somec > 0
• Paris metric. Let x, y ∈ R, the metric is defined as d(x, y) =
|x| + |y|, else.
Example 0.5.2. • ` p space. Let p ≥ 1 be a fixed real number. By definition, each element
in the space ` p is a sequence x = (ξi ) = (ξ1 ; ξ2 , 0, 0, 0) of numbers such that the series
1
∑|ξi | p converges; thus ∑|ξi | p < 1, and the metric is defined by d(x, y) = (∑|ξi − ηi | p ) p ,
where y = (ηi ) and ∑|ηi | p < ∞. We prove that ` p is a metric space. First three properties
of metric are trivial. To prove triangle inequality, we take x; y; z ∈ ` p and using triangle
inequality, we get
1
d(x, y) = ∑|ξi − ηi | p p
1
≤ ∑[|ξi − φi | + |φi − ηi |] p p
1 1
≤ ∑|ξi − φi | p p + ∑|φi − ηi | p p using Minkowski’s inequality
= d(x, z) + d(z, y).
Hence ` p is a metric space.
1/2
• Euclidean space Rn . d(x, y) = ∑ni=1 (ξi − ηi )2 in Rn , where x = (ξ1 , ξ2 , ..., ξn ) and
y = (η1 , ..., ηn ).
• Unitary space Cn . The n-dimensional unitary space is the space of all ordered n-tuples
1/2
of complex numbers with metric defined by d(x, y) = ∑ni=1 (ξi − ηi )2 in Cn , where
x = (ξ1 , ξ2 , ..., ξn ) and y = (η1 , ..., ηn ).
13
0.5. METRIC SPACE
• sequence space `∞ . Let X we take the set of all bounded sequences of complex numbers;
that is, every element of X is a sequence x = (ξ1 , ξ2 , ..., ξn ) written as x = (ξi ) such that
for all i = 1; 2; ..., we have |ξi | ≤ C; where C is a real number which may depend on x,
but does not depend on i. We choose the metric defined by d(x, y) = supi∈N |ξi − ηi |.
• Function space C[a, b]. Let X we take the set of all real-valued functions f , g, ... which
are functions of an independent real variable t and are defined and continuous on a given
closed interval I = [a, b]. Choosing the metric defined by d(x, y) = maxt∈[a,b] | f (t)−g(t)|.
Rb
We can also define another metric on same space as d1 (x; y) = a | f (t) − g(t)|dt.
• d(x, y) = |x − y| (yes)
• d(x, y) = |x3 − y3 | (yes) Here d(x, y) ≥ 0.
Now d(x, y) = 0 =⇒ x3 = y3 =⇒ x2m−1 = y2m−1 =⇒ y = x, xω, xω 2 =⇒ x = y only
over R.
d(x, y) = d(y, x).
Consider
14
0.5. METRIC SPACE
d(x,y)
• If (X, d) is a metric space. Consider d1 (x, y) = 1+d(x,y) Is d1 a metric on X? [yes]
d(x,y)
(i) d1 (x, y) = 1+d(x,y) ≥ 0.
d(x,y)
(ii) d1 (x, y) = 1+d(x,y) = 0 =⇒ d(x, y) = 0 iff x = y.
d(x,y) d(y,x)
(iii) d1 (x, y) = 1+d(x,y) = 1+d(y,x) [as d is metric]. so d1 (x, y) = d1 (y, x).
(iv) Now, we have to prove the triangle inequality. Consider
Euclidean Spaces Let Rn = {(x1 , x2 , ..., xn ) : xi ∈ R∀1 ≤ i ≤ n}. This is a vector space over
R, where sum of two elements x ∈ Rn , y ∈ Rn is x + y = (x1 + y1 , ..., xn + yn ) ∈ Rn and the
product of x ∈ Rn with a scalar λ ∈ R is λ x = (λ x1 , ..., λ x2 ) ∈ Rn .
Define scalar product on Rn by (x, y) = x1 y1 q + ... + xn yn ∈ R.
n
p
Define the norm on R by ||x|| = (x, x) = x12 + ... + xn2 .
Define the map d : Rn × Rn → R by
q
d(x, y) = ||x − y|| = (x1 − y1 )2 + ... + (xn − yn )2 .
The metric space (Rn , d) is called the Euclidean space. The metric d is called the Euclidean
metric.
15
0.6. BASIC PROPERTIES
16
0.6. BASIC PROPERTIES
17
0.7. SOME PROPERTIES IN METRIC SPACE
Proof. (a) Assume that xn → x and xn → y and x 6= y. Then d(x, y) > 0. Then we find a positive
integer k, l such that d(xn , x) < ε/2 for n ≥ k and d(xn , y) < ε/2, for n ≥ l. For n ≥ max{k, l},
we have
d(x, y) = 0. Hence, we conclude that it is impossible for a sequence {xn } to converge to two
different points.
(b) If {xn } converges to x, then we know that there is a positive integer k such that d(xn , x) <
ε for n > k.
Choose M = max{ε, d(x1 , x), ..., d(xk , x)}. Then d(xn , x) ≤ M, for n = 1, 2, 3.... So the
sequence is bounded.
(c) Assume that xn → x and yn → y. Consider
Thus |d(x, y) − d(xn , yn )| ≤ d(xn , x) + d(y, yn ). Now, taking lim n → ∞, we have d(xn , yn ) →
d(x, y).
Definition 0.7.3. Let (X, d) be a metric space and {xn } be a sequence of points of X. We
say that {xn } is Cauchy if for every ε > 0 there exists an integer k such that d(xn , xm ) < ε if
n, m ≥ k.
Proof. Take ε = 1. Since {xn } is Cauchy, there exists an integer k such that d(xn , xm ) < 1 if
n, m ≥ k. Take m = k, then d(xn , xk ) < 1 for n ≥ k.
Let t > 1 be such that d(xi , xk ) < t for 1 ≤ i ≤ k − 1. Then xn ∈ B(xk ,t) for all n, so {xn } is
bounded.
18
0.7. SOME PROPERTIES IN METRIC SPACE
Proof. Assume that xn → x. Then for a given ε > 0, there exists an integer k such that d(xn , x) <
ε/2 if n ≥ k. By taking n, m ≥ k,
d(xn , xm ) ≤ d(xn , x) + d(x, xm ) < ε/2 + ε/2 = ε.
So {xn } is a Cauchy sequence.
Proposition 0.7.4. If {xn } is a Cauchy and it contains a convergent subsequence, then {xn }
converges.
Proof. Assume that {xn } is Cauchy and {xnk } is subsequence of {xn } such that xnk → x. Let
ε > 0. Since {xn } is Cauchy, there exists k0 such that d(xn , xnk ) < ε/2 for all n ≥ k0 . Also, since
xnk → x, there exists k00 such that d(xnk , x) < ε/2 for all n ≥ k00 . Set k = max{k0 , k00 }. Then for
n ≥ k,
d(xn , x) ≤ d(xn , xnk ) + d(xnk , x) < ε/2 + ε/2 = ε.
Hence xn → x.
A Cauchy sequence need not converge. For example, consider sn = {1/n} in the metric
space (X, d), where X = (0, 1), d = |x − y|. Here the sequence is Cauchy in (0,1) but does not
converge to any point of interval as sn → 0 ∈
/ X.
We shall prove that {sn } is a Cauchy sequence.
For ε > 0, choose an integer k such that k > 2/ε. Let m, n ≥ k such that |pm | ≤ 1/m < ε/2
and |pn | ≤ 1/n < ε/2. Now, consider
d(pm , pn ) = |pm − pn |
≤ |pm | + |pn |
≤ 1/m + 1/n
< ε/2 + ε/2 = ε, ∀m, n ≥ k.
Hence {sn } is a Cauchy sequence.
Definition 0.7.4. A metric space (X, d) is called complete if every Cauchy sequence {xn } in X
converges to some point of X.
A subset A of X is called complete if A as a metric subspace of (X, d) is complete, that is,
if every Cauchy sequence {xn } in A converges to a point in A.
Theorem 0.7.5. Let (X, d) be a complete metric space and A ⊆ X. Then A is complete iff A is
closed.
Proof. Suppose that A is a complete subspace of X. We claim that A is closed, that is A = A.
We know that A ⊆ A. Let x ∈ A = A ∪ A0 . So x ∈ A or x ∈ A0 . If x ∈ A0 , x is a limit point of
A ao there exists a sequence {pn } in A such that lim pn = x. Since every convergent sequence
is Cauchy, so {pn } is a Cauchy sequence in A and A is complete. Therefore x ∈ A.
Therefore in both the cases x ∈ A. Hence A ⊆ A. Therefore, A = A. Hence A is closed.
Conversely, Let A be a closed subspace of X. We claim that A is complete.
Let {pn } be a Cauchy sequence in A. As A ⊆ X, so {pn } be a Cauchy sequence in X. Since
X is complete, lim pn = p ∈ X. This implies p ∈ A. Since A is closed, A = A. Hence p ∈ A.
Therefore A is complete.
Or
Let A be complete. Then for every x ∈ A, there is a sequence (xn ) which converges to x.
Since (xn ) is Cauchy and A is complete xn → x. Therefore A is closed.
Conversely, Let A be closed and (xn ) be a Cauchy sequence in A. Then xn → x ∈ A, which
implies x ∈ A, so x ∈ A. Hence A is complete.
19
0.7. SOME PROPERTIES IN METRIC SPACE
Corollary 0.7.6. (i) Every closed subset of a complete metric space is complete.
(ii) Every complete subset of a metric space is closed.
Definition 0.7.5. Let (X, d) and (Y, d1 ) be metric spaces and f : X → Y be a mapping. The
function f is said to be continuous at the point x0 ∈ X if the following holds:
For every ε > 0, there exists a δ > 0 such that for all x ∈ X if d(x, x0 ) < δ , then d1 ( f (x), f (x0 )) <
ε.
Theorem 0.7.7. Let f : X → Y be a function from a metric space (X, d) to (Y, d1 ) and x0 ∈ X.
Then f is continuous at x0 iff for every sequence {xn } such that xn → x0 , f (xn ) → f (x0 ).
Also, f is continuous iff for every convergent sequence {xn } in X, limn→∞ f (xn ) =
f (limn→∞ xn ).
Proof. Suppose that f is continuous at x0 and xn → x0 . We have to prove that f (xn ) → f (x0 ).
Let ε > 0 be given. By definition of continuity at x0 , there exists δ > 0 such that for all x ∈ X,
if d(x, x0 ) < δ , then d1 ( f (x), f (x0 )) < ε.
Since xn → x0 , there exists an integer k such that for all n ≥ k, d(xn , x0 ) < δ .
Hence d1 ( f (xn ), f (x0 )) < ε, for all n ≥ k, that is, f (xn ) → f (x0 ).
Conversely, assume that f is not continuous at x0 . Then there exists some ε > 0 such that for
every δ > 0 there exists a point x ∈ X satisfying d(x, x0 ) < δ but d1 ( f (x), f (x0 )) ≥ ε. For each
n, take δ = 1/n, n=1,2,... and choose xn so that d(xn , x0 ) < 1/n but d1 ( f (xn ), f (x0 )) ≥ ε. Hence
xn → x0 but { f (xn )} does not converge to f (x0 ), which is not true as per our assumption.
20
0.8. NORMED VECTOR SPACE
21
0.8. NORMED VECTOR SPACE
kx + yk p = ∑ |xi + yi| p
i=1
!1 !1
n p n p
≤ ∑ |xi| p + ∑ |yi| p
i=1 i=1
= kxk p + kyk p .
3. Let X = B[a, b] be the set of all bounded real-valued functions on [a, b]. For each x ∈ X,
define kxk∞ = supa≤t≤b |x(t)|. Then (X, k.k∞ ) is a normed linear space.
N1 and N2 are simple to prove. Now, we prove N3, For any t ∈ [a, b] and x, y ∈ X, we
have
|x(t) + y(t)| ≤ |x(t)| + |y(t)| ≤ sup |x(t)| + sup |y(t)| = kxk∞ + kyk∞ .
a≤t≤b a≤t≤b
Since this is true for all t ∈ [a, b], we have that kx + yk∞ = supa≤t≤b |x(t) + y(t)| ≤ kxk∞ +
kyk∞ .
R 1
b 2 2
4. Let X = C[a, b]. For each x ∈ X, define kxk∞ = supa≤t≤b |x(t) and kxk2 = a |x(t)| dt .
Then (X, k.k2 ) and (X, k.k∞ ) are normed linear spaces.
p |
1 ∈ X, define kxk p = (∑i∈N |xi | ) f rac1p. Then
5. Let X = ` p , 1 ≤ p < ∞. For each x = (xi )∞
(X, k.k p ) and is a normed linear space.
Proposition 0.8.2. Let X be a normed space. Then d(x, y) = ||x − y||, x, y ∈ X defines a metric
on X.
Proof. The axioms M1, M2, M3 are easy to see. Now, for x, y, z ∈ X, we have
d(x, z) ≤ ||x − z||
= ||(x − y) + (y − z)||
≤ ||x − y|| + ||y − z||
= d(x, y) + d(y, z).
Hence it is a metric.
Theorem 0.8.3. (a) If (X, k.k) is a normed linear space, then d(x, y) = kx − yk defines a metric
on X. Such a metric d is said to be induced or generated by the norm k.k. Thus, every normed
linear space is a metric space, and unless otherwise specified, we shall henceforth regard any
normed linear space as a metric space with respect to the metric induced by its norm.
(b) If d is a metric on a linear space X satisfying the properties: For all x; y; z ∈ X and for
all α ∈ F,
i) d(x, y) = d(x + z, y + z) (Translation Invariance)
ii) d(αx, αy) = |α|d(x, y) (Absolute Homogeneity); then kxk = d(x, 0) defines a norm on
X.
22
0.8. NORMED VECTOR SPACE
Proof. (a) We show that d(x, y) = kx − yk defines a metric on X. To that end, let x; y; z ∈ X.
M1. d(x, y) = kx − yk ≥ 0 and d(x, y) = 0 ⇔
M2. kx − yk = 0 ⇔ x = y by N1.
M3. d(x, y) = kx − yk = k(−1)(y − x)k = | − 1| ky − xk = ky − xk = d(y, x)
M4.
It is clear from Theorem 0.8.3, that a metric d on a linear space X is induced by a norm on
X if and only if d is translation-invariant and positive homogeneous.
Metric for Rn : We define a norm of a vector n n 2 21
p Xn = (x1 , x2 , ..., xn ) ∈ R as ||X|| = (∑i=1 xi ) .
The metric induce by this norm is d(X,Y ) = ∑i=1 (xi − yi )2 .
1
For p ≥ 1 and X = (x1 , x2 , ..., xn ). Let ||X|| p = (∑ni=1 xi2 ) p . The metric induce by this norm
1
is d p (X,Y ) = (∑ni=1 (xi − yi ) p ) p .
For p = ∞, we have ∞ norm on Rn ||X||∞ = max{|xi | : 1 ≤ i ≤ n}. The metric induce by
this norm is d∞ (X,Y ) = max{|xi − yi | : 1 ≤ i ≤ n}.
Diameter If (X, d) is a metric space then for any subset A of X, Diameter of A is denoted
by dia(A) = d(A) = supx,y∈A d(x, y).
d({a}) = 0, that is diameter of singleton set is zero.
Diameter of a set containing more than one element will be positive.
A metric space is said to be bounded if diameter of X is finite.
|x−y|
Example 0.8.4. Consider metric space (R, d), where (a) d(x, y) = |x − y| (b) d(x, y) = 1+|x−y|
What is the diameter of metric space (R, d)? Is (R, d) bounded?
(a) d(x, y) = |x − y|. Here dia(R) = ∞, because dia(R) = sup d(x, y) = sup[0, ∞) = ∞.
Hence metric (R, d) is not bounded.
|x−y|
(b) d(x, y) = 1+|x−y| < 1. dia(R) = sup d(x, y) = sup[0, 1) = 1. Hence metric (R, d) is
bounded.
23
0.9. CONVEX SET
In Euclidean space the cube, circle, sphere, triangular area in Euclidean 2-space, Euclidean
plane and line segments are convex sets.
Thus a set C is convex in X if for all scalars α in 0 ≤ α ≤ 1, αC +(1−α)C ⊂ C, equivalently
for any two reals α, β with 0 ≤ α, β < 1, αC + βC ⊂ C.
Theorem 0.9.1. Let x, y, z be three distinct points in the real linear space L such that z ∈ `x,y .
Then, one of these points belongs to the open segment determined by the remaining two points.
Proof. Since z ∈ `x,y , we have z = (1 − a)x + ay for some a ∈ R.
y
a>1
u
x
u 0<a<1
a<0
1
We have a 6∈ {0, 1} because the points x, y, z are distinct. If a > 1 we have y = a−1
a x + a z,
a−1 1
so y ∈ (x, z) because a , a ∈ (0, 1). If 0 < a < 1 we have z ∈ (x, y). Finally, if a < 0, since
a −a
x = 1 + 1−a z + 1−a y, we have x ∈ (z, y).
24
0.9. CONVEX SET
Theorem 0.9.2. Let L be a real linear space. A subset C of L is convex if and only if any convex
combination of elements of C belongs to C.
or
Let L be a linear space and C ⊂ L be a convex set. If x1 , x2 , · · · , xn ∈ C then all elements of
the form α1 x1 + α2 x2 + · · · + αn xn ∈ C where αi ≥ 0 and α1 + α2 + · · · + αn = 1.
Proof. The sufficiency of this condition is immediate. To prove its necessity consider x1 , . . . , xk ∈
C and the convex combination
y = a1 x1 + · · · + ak xk .
We prove by induction on k ≥ 1 that y ∈ C. The base case, m = 1 is immediate since in this
case y = a1 x1 and a1 = 1.
For the inductive step, suppose that the statement holds for k and let y be given by y =
a1 x1 + · · · + ak xk + ak+1 xk+1 , where a1 + · · · + ak + ak+1 = 1, ai ≥ 0 and xi ∈ C for 1 ≤ i ≤ k + 1.
We have
k
ai
y = (1 − ak+1 ) ∑ xi + ak+1 xk+1 .
i=1 1 − ak+1
Theorem 0.9.3. Arbitrary intersection of any number of convex sets in a linear space is a
convex set.
Proof. Let M = {Ci : i ∈ I} be a collection of convex sets and let C = M. Suppose that
T
Theorem 0.9.4. Let C be a convex subset of a real linear space L. If r1 , r2 ∈ R≥0 , then we have
Proof. If at least one of r1 , r2 is 0 the equality obviously holds; therefore, assume that both r1
and r2 are positive.
Let z ∈ r1C + r2C. There exists x, y ∈ C such that z = r1 x + r2 y, and therefore,
r1 r1
z = (r1 + r2 ) x+ y .
r1 + r2 r1 + r2
Theorem 0.9.5. If A and B are two convex sets in a linear space X then A + B is also a convex
set.
25
0.9. CONVEX SET
x = α1 x1 + α2 x2 + · · · + αn xn ,
26
0.9. CONVEX SET
x = α1 x1 + α2 x2 + · · · + αm xm + αm+1 xm+1
= β β1 x1 + β β2 x2 + · · · + β βm xm + (1 − β )xm+1
= β (β1 x1 + β2 x2 + · · · + βm xm ) + (1 − β )xm+1 ∈ T ⊂ W.
So x = α1 x1 + α2 x2 + · · · + αm xm + αm+1 xm+1 ∈ W .
Thus the result is true by the Principle of Mathematical Induction, i.e. every elements of T
is an element of W with S ⊂ W . As W is arbitrary it follows that T ⊂ ∩W , where S ⊂ W , that
is T ⊂ conv hull(S).
Therefore T = conv hull(S). Hence the result.
27
0.10. QUOTIENT SPACE
Proposition 0.10.1. Let M be a linear subspace of a linear space X over F. For x, y ∈ X and
α ∈ F, define the operations [x + y] = [x] + [y] and [αx] = α[x]. Then X/M is a linear space
with respect to these operations.
Proposition 0.10.3. Let M be a closed linear subspace of a normed linear space X over F.
The quotient space X/M is a normed linear space with respect to the norm k[x]k = infy∈[x] kyk,
where [x] ∈ X/M.
28
0.10. QUOTIENT SPACE
The norm on X/M as defined in Proposition 0.10.3 is called the quotient norm on X/M.
Theorem 0.10.4. If X is a normed space, and Y is a closed normed linear subspace of X, then
X/Y is a normed space under the norm kx +Y k = infz∈x+Y kzk.
Proposition 0.10.5. (a) Show that k.k is continuous, that is, k.k : X → R defined by x 7→ kxk is
a continuous mapping.
(b) Show that the non trivial mapping f (x) = Ax from Rn to Rm (with Euclidean norms)
defined by the m × n matrix A = (ai j ) is uniformly continuous.
29
0.10. QUOTIENT SPACE
30
0.11. COMPLETENESS PROPERTY IN NORMED LINEAR SPACE
Proposition 0.11.2. Let (X, k.k) be a normed linear space over F. A Cauchy sequence in X
which has a convergent subsequence is convergent.
Proof. Let (xn ) be a Cauchy sequence in X and {xnk } its subsequence which converges to x ∈ X.
Suppose that for a given ε > 0, there is a natural numbers N1 , N2 such that kxn − xm k < ε2 for
all n, m ≥ N1 and kxnk − xk < ε2 for all k ≥ N2 .
Let N = max{N1 , N2 }. If k ≥ N, then since nk ≥ k, kxk − xk ≤ kxn − xnk k + kxnk − xk <
ε ε
2 + 2 = ε. Hence xn → x.
Definition 0.11.2. A metric space (X, d) is said to be complete if every Cauchy sequence in X
converges in X.
31
0.11. COMPLETENESS PROPERTY IN NORMED LINEAR SPACE
Definition 0.11.3. A normed linear space that is complete with respect to the metric induced
by the norm is called a Banach space.
Theorem 0.11.3. Let (X, k.k) be a Banach space and M be a linear subspace of X. Then M is
complete if and only if M is closed in X.
Proof. Assume that M is complete. We have to show that M is closed. To that end, let x ∈ M.
Then there is a sequence (xn ) in M such that kxn − xk → 0. Since (xn )converges, it is Cauchy.
Completeness of M guarantees the existence of an element y ∈ M such that kxn − yk → 0. By
uniqueness of limits, x = y. Hence x ∈ M and, consequently, M is closed.
Assume that M is closed. We show that M is complete. Let (xn ) be a Cauchy sequence in
M. Then (xn ) is a Cauchy sequence in X. Since X is complete, there is an element x ∈ X such
that kxn − xk → 0. But then x ∈ M since M is closed. Hence M is complete.
32
0.11. COMPLETENESS PROPERTY IN NORMED LINEAR SPACE
(m) (m)
(b) Let (xm ) be a Cauchy sequence in `∞ , where xm = (ξ1 , ξ2 , ...) ∈ `∞ . Since the metric
defined on `∞ is given by d(x, y) = supi |ξi − ηi |, where x = (ξi ) and y = ηi . For a given ε > 0,
(m) (n)
there is an N such that for all m, n ≥ N, we have d(xm , xn ) = supi |ξi − ξi | < ε.
(m) (n)
Therefore, for every fixed j, and m, n ≥ N we have |ξi − ξi | < ε. Hence for every fixed
(1) (2)
j, the sequence (ξ j , ξ j , ...) is a Cauchy sequence in R. As R is complete, so it converges, say
(m) (m) (m) (m)
ξj → ξ j as m → ∞, we have |ξ j − ξ j | < ε, whenever m > N. Since xm = (ξ1 , ξ2 , ...) ∈
(m)
`∞ , there is a real number km such that |ξ j | ≤ km for all j. Hence by triangle inequality
(m) (m)
|ξ j | ≤ |ξ j − ξ j | + |ξ j | ≤ ε + km for m > N. This inequality holds for every j, and the right
hand side does not involve j. Hence (ξ j ) is a bounded sequence of numbers. This implies
(m)
that x = (ξ j ) ∈ `∞ . Also, we obtain d(xm , x) = supi |ξ j − ξ j | ≤ ε, for m > N. This show that
xm → x. Hence `∞ is complete.
(c) Let (xn ) be a a Cauchy sequence in ` p . Suppose that xn = (xn(1) , xn(2) ...). Then for given
1
p p < ε for all m, n ≥ N.
ε > 0, there exists N(ε) ∈ N such that d(xn , xm ) = ∑∞ |x
i=1 n(i) − x m(i) |
For each fixed index i, we have |xn(i) − xm(i) | < ε for all m, n ≥ N, that is for each fixed
index i, (xn(i) )∞ 1 is a Cauchy sequence in R. Since R is complete, there exists x(i) ∈ R such that
xn(i) → x(i) as n → ∞, that is for each fixed index i, (xn(i) )∞ 1 is a Cauchy sequence in R. Since
R is complete, there exists x(i) ∈ R such that xn(i) → x(i) as n → ∞. Define x = (x1 , x2 , ...).
Now, we show that x ∈ ` p and xn → x. For each k ∈ N, ∑ki=1 |xn(i) − xm(i) | p ≤ [d(xn , xm )] p =
p < ε p , that is k |x
∑i=1 n(i) − xm(i) | p < ε p , for all k = 1, 2, .... Keep k and n ≥
∑∞ i=1 |xn(i) − xm(i) |
N fixed and let m → ∞, we have ∑ki=1 |xn(i) − xi | p < ε p . Now, letting k → ∞, then for all n ≥ N,
we have ∑∞ p p
i=1 |xn(i) − xm(i) | < ε , which means that xn − x ∈ ` p . Since xn ∈ ell p , we have to
show that x = (x − xn ) + xn ∈ ` p . Consider |xi | = |xi − xn(i) + xn(i) |, using Minkowski inequality,
we have (∑i |xi | p )(1/p) = (∑i |xi − xn(i) + xn(i) | p )(1/p) ≤ (∑i |xi − xn(i) | p )1/p + (∑i xn(i) | p )(1/p) ).
Now, both the series in right converges so series in left converges. Hence x = (x − xn ) + xn ∈ ` p
Also, xn → x. Hence the result.
1
Example 0.11.5. 1. The space Rn equipped with k.k p defined as kxk2 = ∑i∈N |xi |2 2 , is a
Banach space.
1
(In metric d(x, y) = ∑ni=1 (χi − ηi )2 2 , where 1 ≤ p < ∞, x = (χi ) and y = (ηi )). It
is easy to see that kxk2 (respectively, d(x, y)) defined above is a normed linear space
(respectively, metric space).
Now, we have to show that it is complete. Let (xn )n∈N be a sequence in Rn . Then we have
kxk − xm k < ε for all m, k ≥ N, that is, ∑ni=1 | kxki − xmi k |2 < ε 2 for all m, k ≥ N. Thus
it follows that for every i ∈ {1, 2, ..., n}, |xki − xmi | < ε for all m, k ≥ N, that is sequence
(xmi ) is a Cauchy sequence in R and consequently it is convergent. Let xi = lim xki . Then
k→∞
x = (x1 , ..., xn ) ∈ Rn . Now, take k → ∞, we have ∑ni=1 | kxi − xmi k |2 < ε 2 for all m ≥ N,
that is kx − xm k2 ≤ ε for all m ≥ N, and so x = lim xm in the normed linear space.
m→∞
2. Let 1 ≤ p < ∞. The sequence space ` p is a Banach space. Because of the importance of
this space, we give a detailed proof of its completeness.
The classical sequence space ` p is complete.
Proof. Let (xn ) be a a Cauchy sequence in ` p . Suppose that xn = (xn(1) , xn(2) ...). Then
1
p p <ε
for given ε > 0, there exists N(ε) ∈ N such that kxn − xm k p = ∑∞ |x
i=1 n(i) − xm(i) |
33
0.11. COMPLETENESS PROPERTY IN NORMED LINEAR SPACE
for all m, n ≥ N.
For each fixed index i, we have |xn(i) − xm(i) | < ε for all m, n ≥ N, that is for each
fixed index i, (xn(i) )∞ 1 is a Cauchy sequence in R. Since R is complete, there exists
x(i) ∈ R such that xn(i) → x(i) as n → ∞, that is for each fixed index i, (xn(i) )∞ 1 is a
Cauchy sequence in R. Since R is complete, there exists x(i) ∈ R such that xn(i) →
x(i) as n → ∞. Define x = (x1 , x2 , ...). Now, we show that x ∈ ` p and xn → x. For
1 1
each k ∈ N, ∑ki=1 |xn(i) − xm(i) | p p ≤ kxn − xm k p = ∑∞ i=1 |xn(i) − xm(i) |
p p < ε, that is
∑ki=1 |xn(i) − xm(i) | p < ε p , for all k = 1, 2, .... Keep k and n ≥ N fixed and let m → ∞, we
have ∑ki=1 |xn(i) − xi | p < ε p . Now, letting k → ∞, then for all n ≥ N, we have ∑∞ i=1 |xn(i) −
xm(i) | p < ε p , which means that xn − x ∈ ` p . Since xn ∈ ` p , we have x = (x − xn ) + xn ∈ ` p .
Also, xn → x. Hence the result.
3. Space `∞ is complete.
(m) (m)
Proof. Let (xm ) be a Cauchy sequence in `∞ , where xm = (ξ1 , ξ2 , ...) ∈ `∞ . Since the
metric defined on `∞ is given by d(x, y) = supi |ξi − ηi |, where x = (ξi ) and y = ηi . For
(m)
a given ε > 0, there is an N such that for all m, n ≥ N, we have d(xm , xn ) = supi |ξi −
(m)
ξi | < ε.
Therefore, for every fixed j, and m, n ≥ N we have
(m) (m)
|ξi − ξi | < ε.
(1) (2)
Hence for every fixed j, the sequence (ξ j , ξ j , ...) is a Cauchy sequence in R. As R is
(m) (m)
complete, so it converges, say ξ j → ξ j as m → ∞, we have |ξ j − ξ j | < ε, whenever
(m) (m) (m)
m > N. Since xm = (ξ1 , ξ2 , ...) ∈ `∞ , there is a real number km such that |ξ j | ≤ km
(m) (m)
for all j. Hence by triangle inequality |ξ j | ≤ |ξ j − ξ j | + |ξ j | ≤ ε + km for m > N.
This inequality holds for every j, and the right hand side does not involve j. Hence (ξ j )(
is a bounded sequence of numbers. This implies that x = (ξ j ) ∈ `∞ . Also, we obtain
(m)
d(xm , x) = supi |ξ j − ξ j | ≤ ε, for m > N. This show that xm → x. Hence `∞ is complete.
This shows that the sequence x = (ξ j ) is convergent. Hence x ∈ c. Since x ∈ c, this proves
that closedness of c in `∞ and it is complete.
Some examples of not complete space.
34
0.11. COMPLETENESS PROPERTY IN NORMED LINEAR SPACE
R1
Figure 1: This is the graph of C[a, b] induced with d(x, y) = 0 |x(t) − y(t)|dt
Example 0.11.6. 1. Let X be the set of all continuous real-valued functions on I = [0, 1] and
d(x, y) = 01 |x(t) − y(t)|dt. Then (X, d) is not complete.
R
Proof.
The functions
0,t ∈ [0, 1/2]
xm (t) = m(x − 12 ), 12 < t < am = 1/2 + 1/m
1,t ∈ [a , 1],
m
form a Cauchy sequence because d(xm , xn ) is the area of the triangle in Figure 1 and
for every given ε > 0, d(xm , xn ), when m, n > 1/ε. Now, we’ll show that this Cauchy
sequence does not converge in X. For every x ∈ X, we have
Z 1
d(xm , x) = |xm (t) − x(t)|dt
0
Z 1/2 Z am Z 1
= |x(t)|dt + |xm (t) − x(t)| + |1 − x(t)|dt.
0 1/2 am
Each the integrands are nonnegative, so is each integral on the right. Hence d(xm , x) → 0
would imply that each integral approaches zero and, since x(t) is continuous, we should
have (
0,t ∈ [0, 1/2)
x(t) =
1,t ∈ (1/2, 1]
But this is impossible for a continuous function. Hence (xm ) does not converge, that is,
does not have a limit in X. This proves that X is not complete.
Example 0.11.7. Show that (C[a, b], d∞ ) is a complete metric space, where d∞ = sup |x(t) −
y(t)|, t ∈ [a, b].
Proposition 0.11.8. Show that the set of all real numbers constitutes an incomplete metric
space if we choose the metric
35
0.11. COMPLETENESS PROPERTY IN NORMED LINEAR SPACE
Proof. Consider a sequence (xn ) on real line, where xn = n. Then (xn ) is Cauchy but not
convergent. We know that for a given ε > 0, there exists an integer N such that for n > N,
On the other hand, (xn ) is not a convergent sequence. If not there exists x ∈ R such that
d(xn , x) → 0 as n → ∞. But (using |x − y| ≥ ||x| − |y||)
So tan−1 x = π/2. But this a contradiction because for any x ∈ R, tan−1 x < π/2.
36
0.12. CONVERGENCE OF SERIES IN BANACH SPACE
Theorem 0.12.1. A normed linear space (X, k.k) is a Banach space if and only if every abso-
lutely convergent series in X is convergent.
Proof. Let X be a Banach space and ∑∞j=1 x j < ∞. Now, we have to show that ∑∞j=1 x j
converges. Suppose for a given ε > 0, and for each n ∈ N, let sn = ∑nj=1 x j . Let K be a positive
integer such that ∑∞j=K+1 x j < ε. Then for all m > n > K, we have
m n m
ksm − sn k = ∑ xj − ∑ xj = ∑ xj
j=1 j=1 j=n+1
m
≤ ∑ xj
j=n+1
∞
.
≤ ∑ xj
j=n+1
∞
≤ ∑ x j < ε.
j=K+1
Hence the sequence (sn ) of partial sums form a Cauchy sequence in X. SInce X is complete,
the sequence (sn ) converges to some element s ∈ X, that is, the series ∑∞j=1 x j converges.
Conversely, assume that (X, k.k) is a normed linear space in which every absolutely con-
vergent series converges. We have to show that X is complete. Let (xn ) be a Cauchy se-
quence in X. Then there exists n1 ∈ N such that kxn1 − xm k < 12 , where m > n1 . Simi-
larly, there is an n2 ∈ N with n2 > n1 such that kxn2 − xm k < 212 whenever m > n2 . Contin-
uing in this way, we get natural numbers n1 < n2 < ... such that kxnk − xm k < 21k whenever
m > nk . In particular, we have that for each k ∈ N, xnk+1 − xnk < 2−k . For each k ∈ N, let
yk = xnk+1 − xnk . Then ∑nk=1 kyk k = ∑nk=1 xnk+1 − xnk < ∑nk=1 2−k . Hence ∑nk=1 kyk k < ∞, that
is series ∑nk=1 yk is absolutely convergent and hence by our assumption, the series ∑nk=1 yk
j
is convergent in X, that is, there is an x ∈ X such that s j = ∑k=1 yk → s. It follows that
j j
s j = ∑k=1 yk = ∑k=1 (xnk+1 − xnk ) = xn j+1 − xn1 → s as j → ∞. Hence xn j+1 → s + xn1 . Thus the
subsequence (xnk ) of (xn ) converges in X. But if a Cauchy sequence has a convergent subse-
quence, then the sequence itself also converges (to the same limit as the subsequence). It thus
follows that the sequence (xn ) also converges in X. Hence X is complete.
Theorem 0.12.2. Let a, b ∈ R and a < b. The space (C[a, b], k.k∞ ) is a Banach space.
Proof. It is clear that linear combinations of continuous functions are continuous, so that
C[a, b] is a vector space. Define k f k∞ = supx∈[a,b] | f (x)|, f ∈ C[a, b]. Then C[a, b] is a normed
linear space.
37
0.12. CONVERGENCE OF SERIES IN BANACH SPACE
Now, we have to show the completeness. Let ( fn ) be a Cauchy sequence in C[a, b]. Suppose
for a given ε > 0, there exists a N ∈ N such that for all x ∈ [a, b], we have
Hence f is continuous and it belongs to C[a, b].Hence k f − fn k∞ < ε, for all n ≥ N and so fn
converges to f in the normed space C[a, b].
38
0.13. CANTOR INTERSECTION THEOREM
Theorem 0.13.3. Cantor’s Intersection Theorem Let (X, d) be a metric space and (Fn ) be a
nested sequence of non-empty closed subsets of X such that diam(Fn ) → 0. Then X is complete
iff ∩∞
n=1 Fn consists of exactly one point.
Proof. Let X be complete. For each n, we choose xn ∈ Fn . Since diam(Fn ) → 0, for every
ε > 0, there exists a positive integer m0 such that diam(Fm0 ) < ε. Again since (Fn ) is a nested
sequence, we have for n, m > m0 , Fn , Fm ⊂ Fm0 and hence xn , xm ∈ Fm0 , d(xn , xm ) < ε. Thus
(xn ) is a Cauchy sequence. Since X is complete, xn → x0 for some x0 ∈ X.
Now, we have to show that x0 ∈ ∩∞ n=1 Fn . Assume that m ∈ N, for n > m, we have xn ∈
Fn ⊂ Fm . Since xn → x0 , the sequence (xn ) is in every neighbourhood of x0 . As, xn ∈ Fm , so x0
contains an infinite number of points of Fm . Thus x0 is a limit point of Fm . Since Fm is closed,
x0 ∈ Fm and m is arbitrary, we have x0 ∈ ∩∞ n=1 Fn .
Now, suppose that there is another point x1 (6= x0 ) ∈ ∩∞ n=1 Fn . Therefore, d(x0 , x1 ) < diam(Fn )
for every n. Therefore d(x0 , x1 ) = 0, since diam(Fn ) → 0. Hence x0 = x1 and so ∩∞ n=1 Fn = {x0 }.
Conversely, let ∩∞ F
n=1 n consists of a single point for every nested sequence (F n ) of nonempty
closed subsets of Fn of X such that diam(Fn ) → 0.
39
0.13. CANTOR INTERSECTION THEOREM
Now, we have to prove that X is complete. Let (xn ) be a Cauchy sequence in X. Take
S1 = {x1 , x2 , ...}, S2 = {x2 , x3 , ...} ,..., Sn = {xn , xn+1 , ...}. Since (xn ) is a Cauchy sequence, for
given ε > 0, there exists a positive integer m0 such that n, m > m0 , d(xn , xm ) < ε. It follows
that for n > m0 , diam(Sn ) < ε and consequently, diam(Sn ) → 0.
Also, S1 ⊃ S2 ⊃ ... so that S1 ⊃ S2 ⊃ ... [as A ⊃ B =⇒ A ⊃ B]. Therefore, diam(Sn ) =
diam(Sn ). Hence (Sn ) is a nested sequence of closed subsets of X whose diameter tends to
zero. Then by hypothesis, there exists a unique point x0 ∈ X such that x0 ∈ ∩∞ n=1 Sn . We claim
that (xn ) converges to x0 . Since diam(Sn ) → 0, for given ε > 0, there exists m0 ∈ N such that
diam(Sm0 ) < ε and consequently, for n > m0 , xn , x0 ∈ Sm0 , d(xn , x0 ) < ε. Hence (xn ) converges
to x0 . Thus it is shown that every Cauchy sequence in X converges to a point in X. Hence X is
complete.
40
0.14. SEPARABLE SPACE
e1 = (1; 0; 0; 0; ...)
e2 = (0; 1; 0; 0; ...)
e3 = (0; 0; 1; 0; ...)
.......
If a normed space X has a Schauder basis, then X is separable. Conversely, does every
separable Banach space have a Schauder basis? This is a famous question raised by Banach
himself. Almost all known separable Banach spaces had been shown to possess a Schauder
basis. Nevertheless, the surprising answer to the question is no and answer was given by P.
Enflo (1973) who was able to construct a separable Banach space which has no Schauder
basis.
Definition 0.14.1. A subset M of a metric space X is said to be dense in X if M = X.
X is said to be separable if it has a countable subset which is dense in X.
That is, we can have limit points of A within it and the result can equal to the parent set. One
good example is the set of rationals and the real set. We all know that the set of rationals is not
complete. In Analysis, real numbers can be constructed using Dedekind cuts or the addition of
limits to every Cauchy sequence. That is, Q = R or that the set of rationals are dense in the set
of reals. The complex plane, too, can be separated from the irrational real and imaginary parts
against the rational ones. This has importance in the theory of operators.
More technically, if M is dense in X, then for each x ∈ X and for every ε > 0, B(x0 ; ε)
will contain points of M; or, in other words, in this case there is no point x0 ∈ X which has
a neighbourhood that does not contain points of M. This is the direct consequence of the
definition of a limit point.
Examples [1] The real line R is separable since the set Q of rational numbers is a countable
dense subset of R.
[2] The complex plane C is separable since the set of all complex numbers with rational
real and imaginary parts is a countable dense subset of C.
41
0.14. SEPARABLE SPACE
Proof. We will proceed as follows: we will first construct a countable subset then basing our
argument on the fact that Q is dense in R, construct limits for every sequence of elements of
` p which will be limit points of sequences in the constructed subset.
Let M be the set of all sequences x of the form
x = (χ1 ; χ2 ; . . . ; χn ; 0; 0; . . .),
where n is any integer. Now, we can assume that χi ∈ Q for all i since we are only discussing
real or complex sequences. Since Q is countable, Qn is therefore countable, leading us to a
countable M. That justifies one part of the definition. To prove that M = ` p . Take y = (ηi ) ∈ ` p .
Since this is convergent, we have ∑∞ p p
k=n+1 |ηk | < ε /2. Since Q = R, for each ηk there is
rational χk close to it. Hence we can find a x ∈ M such that ∑nk=1 |χk − ηk | p < ε p /2. Since
1
p p p p n
d(x, y) = (∑∞ ∞
k=1 |χk − ηk | ) . We therefore have [d(x, y)] = ∑k=n+1 |ηk − 0| + ∑k=1 |χk −
ηk | p < ε p /2 + ε p /2 = ε p or that d(x, y) < ε. That is, every sequence x will have a limit point
y. Hence, M = ` p .
or
Take M to be the set of all sequences with rational entries such that all but a finite number of
the entries are zero. (If the entries are complex, take for M the set of finitely nonzero sequences
with rational real and imaginary parts.) It is clear that M is countable. We show that M is dense
in ` p . Let ε > 0 and x = (xn ) ∈ ` p . Then there is an N such that
∞
∑ |xk | p < ε/2.
k=N+1
Now, for each 1 ≤ k ≤ N, there is a rational number qk such that |xk − qk | p < ε/(2N). Set
q = (q1 , q2 , . . . , qN , 0, 0, . . .). Then q ∈ M and
N ∞
kx − qk pp = ∑ |xk − qk | p + ∑ |xk | p < ε.
k=1 k=N+1
Hence M is dense in ` p .
This should in no way mean that every collection of convergent sequences forms a separable
set.
Theorem 0.14.2. A normed linear space X is separable if and only if it contains a countable
set B such that lin(B) = X.
Proof. Assume that X is separable and let A be a countable dense subset of X. Since the linear
hull of A, lin(A), contains A and A is dense in X , we have that lin(A) is dense in X , that is,
lin(A) = X.
Conversely, assume that X contains a countable set B such that lin(B) = X. Let B = {xn :
n ∈ N}.
Assume first that F = R, and put
n
C = { ∑ λ j x j : λ j ∈ Q; j = 1; 2; . . . ; n; n ∈ N}.
j=1
We first show that C is a countable subset of X . The set Q × B is countable and conse-
quently, the family F of all finite subsets of Q × B is also countable. The mapping
n
{(λ1 , x1 ), . . . , (λn , xn )} 7→ ∑ λ jx j
j=1
42
0.14. SEPARABLE SPACE
Clearly, en ∈ ` p . Let ε > 0 and x = (xn ) ∈ ` p . Then there is a natural number N such that
p
∑∞
k=n+1 |xk | < ε, for all n ≥ N.
Now, if n ≥ N, then Hence lin({en : n ∈ N}) = ` p . Of course, the set {en : n ∈ N} is
countable.
Proof. Let y = (ηi ), where ηi = 0, 1. There are uncountably many y. If we put small balls
with radius 13 at y, they will not intersect. It follows that if M ⊂ `∞ is dense in `∞ , then M is
uncountable. Therefore `∞ is not separable.
43
0.15. FINITE DIMENSIONAL NORMED SPACE
Hence it suffices to prove the existence of a c > 0 such that (0.15.2) holds for every n-tuple of
scalars β1 , ..., βn with ∑nj=1 |β j | = 1.
Suppose that this is false. Then there exists a sequence (ym ) of vectors
n
(m) (m) (m)
ym = β1 x1 + ... + βn xn = ∑ |β j |xi ,
j=1
(m)
where ∑nj=1 |β j | = 1 such that kym k → 0 as m → ∞.
(m) (m) (m)
Since ∑nj=1 |β j | = 1, we have |β j | ≤ 1. Hence for each fixed j the sequence (β j ) =
(1) (2) (m)
(β j , β j , ...) is bounded. Consequently, by the Bolzano-Weierstrass theorem, (β1 ) has a
convergent subsequence. Let β1 denote the limit of that subsequence, and let
n
(m) (m)
(y1,m ) = γ1 x1 + ∑ |βi |xi
i=2
(m)
denote the corresponding subsequence of (ym ) such that γ1 → β1 .
(m) (m) (m)
By the same argument, (y1,m ) has a subsequence (y2,m ) = γ1 x1 + γ2 x2 + ∑nj=3 |β j |x j
(m)
for which the corresponding subsequence of scalars β2 converges; let β2 denote the limit.
Continuing in this way, after n steps we obtain a subsequence (yn,m ) = (yn,1 , yn,2 , ...) of
(m) (m) (m)
(ym ) whose terms are of the form yn,m = ∑nj=1 γ j x j , where ∑nj=1 |γ j | = 1, with scalars γ j
(m)
satisfying γ j → β j as m → ∞. Hence as m → ∞, we have yn,m → y = ∑nj=1 β j x j , where
∑nj=1 |β j | = 1, so that not all β j can be zero. Since {x1 , ...xn } is a linearly independent set,
we thus have y 6= 0. On the other hand, yn,m → y implies kyn,m k → kyk, by the continuity of
the norm. Since kym k → 0 by assumption and (yn,m ) is a subsequence of (ym ), we must have
kyn,m k → 0. Hence kyk = 0, so that y = 0. This contradicts y 6= 0. Hence we get the result.
44
0.15. FINITE DIMENSIONAL NORMED SPACE
Proof. We consider an arbitrary Cauchy sequence (ym ) in Y and show that it is convergent in
Y ; the limit will be denoted by y. Let dimY = n and {e1 , ..., en } any basis for Y . Then each ym
(m)
has a unique representation of the form ym = ∑nj=1 α j e j .
Since (ym ) is a Cauchy sequence, for every ε > 0 there is an N such that kym − yr k < ε
when m, r > N. From this and Lemma 0.15.1 we have for some c > 0,
n n
(m) (r) (m) (r)
ε > kym − yr k = ∑ (α j − α j )e j ≥ c ∑ |α j − α j |,
j=1 j=1
(m) (1)
m, r > N. This shows that each of the n sequences (α j ) = (α j , ...), j = 1, ..., n is Cauchy in
(m)
R or C. Hence it converges, let α j → α j denote the limit. Using these n limits α1 , ...αn , we
define y = α1 e1 + ... + αn en . Clearly, y ∈ Y . Furthermore,
n n
(m) (r) (m) (r)
kym − yk = ∑ (α j − α j )e j ≤ ∑ |α j −αj | ej .
j=1 j=1
(m)
As α j → α j , kym − yk → 0, that is ym → y. This shows that (ym ) is convergent in Y . Since
(ym ) was an arbitrary Cauchy sequence in Y , this proves that Y is complete.
Theorem 0.15.3. A subspace M of a complete metric space X is itself complete if and only if
the set M is closed in X.
Using above two theorems, we have
Theorem 0.15.4. Every finite dimensional subspace Y of a normed space X is closed in X.
Note that infinite dimensional subspaces need not be closed or complete.
Example 0.15.5. Let X = C[0, 1], k.k = maxt∈[0,1] |x(t)| and Y = span{x0 , x1 , ...}, where x j (t) =
t j , so that Y is the set of all polynomials, that is Y = span{1,t,t 2 , ...}. Y ⊂ C[0, 1]. Y is not
closed as yn (t) = 1 + t + t 2 /(2!) + ... + t n /(n!) → et ∈
/ Y . Hence Y is not complete.
45
0.16. EQUIVALENT NORMS
for all x ∈ X.
Example 0.16.1. Let X = R2 , x = (ξ1 , ξ2 ) ∈ X and define
and q
kxk2 = |ξ1 |2 + |ξ2 |2 .
Here [By Holder’s inequaity]
2 q
2 1/2 2 1/2
kxk1 = ∑ (1)|ξi | ≤ (∑ 1 ) (∑|ξi | ) = 2 kxk2
i=1
or √1 kxk≤ kxk2 .
2 1
p
Also, kxk2 = |ξ1 |2 + |ξ2 |2 ≤ |ξ1 | + |ξ2 | = kxk1 . Hence
1
√ kxk1 ≤ kxk2 ≤ 1 kxk1 .
2
Therefore kxk1 and kxk2 are equivalent.
Example 0.16.2. Let X = Rn , for each x = (x1 , ..., xn ) ∈ X, let
!1
n n 2
46
0.16. EQUIVALENT NORMS
Lemma 0.16.3. Let {x1 , ..., xn } be a linearly independent set of vectors in a normed space X
(of any dimension). Then there is a number c > 0 such that for every choice of scalars α1 , ...αn
we have
kα1 x1 + ... + αn xn k ≥ c(|α1 | + ...|αn |), (c > 0). (0.16.1)
Theorem 0.16.4. On a finite dimensional normed space X every two norms k.k1 and k.k2 are
equivalent.
Proof. Let n = dim X, and {e1 , e2 , ..., en } be a basis for X with kei k = 1. Assume that x ∈ X
represented as x = ∑ni=1 αi ei . Then by lemma there exists c > 0 such that kxk1 ≥ c ∑ni=1 |αi |.
Also, using triangle inequality, we have kxk2 ≤ ∑ni=1 |αi | kei k2 ≤ k ∑ni=1 |αi |, where k =
max{kei k2 }.
Combining these inequalities, we get kxk2 ≤ (k/c)c ∑ni=1 |αi | ≤ (k/c) kxk1 = (1/a) kxk1 ,
where a = c/k. Hence a kxk2 ≤ kxk1 .
By interchanging the roles of k.k1 and k.k2 and proceeding as above we get the other
inequality. Hence the result.
Proposition 0.16.5. Show that two norms k.k1 and k.k2 on Rn satisfy √1 kxk ≤ kxk2 ≤ kxk1 .
n 1
Proof. Since kxk21 = (∑ni=1 |ξi |)2 ≥ ∑ni=1 |ξi |2 = kxk1 , thus kxk1 ≥ kxk2 .
1/2
2 1/2
1 1
n n
Also, √n kxk1 = ∑i=1 √n |ξi | ≤ ∑i=1 |ξi | 2
∑i=1 √1n
n
= kxk2 .
Therefore, √1 kxk ≤ kxk2 ≤ kxk1 .
n 1
47
0.17. COMPACTNESS
0.17 Compactness
Definition 0.17.1. A metric space X is said to be compact if every sequence in X has a con-
vergent subsequence. A subset M of X is said to be compact if M is compact considered as a
subspace of X, that is, if every sequence in M has a convergent subsequence whose limit is an
element of M.
Theorem 0.17.2 (Bolzano-Weistrass Theorem). Every real, bounded sequence has at least one
convergent subsequence.
Theorem 0.17.3. In a finite dimensional normed space X, any subset M ⊂ X is compact if and
only if M is closed and bounded.
Lemma 0.17.4. F. Riesz’s Lemma. Let Y and Z be subspaces of a normed space X (of any
dimension), and suppose that Y is closed and is a proper subset of Z. Then for every real
number θ in the interval (0, 1) there is a z ∈ Z such that kzk = 1, kz − yk ≥ θ for all y ∈ Y .
48
0.17. COMPACTNESS
Proof. We consider any v ∈ Z −Y and denote its distance from Y by a, that is a = inf kv − yk
y∈Y
(see figure).
Z X
v Y
y0
a = inf kv − yk
y∈Y
Clearly, a > 0 since Y is closed. We now take any θ ∈ (0, 1). By the definition of an
infimum there is a y0 ∈ Y such that
a ≤ kv − y0 k ≤ a/θ , (0.17.1)
a/θ > a since 0 < θ < 1. Let z = c(v − y0 ), where c = 1/ kv − y0 k. Then kzk = 1, and we
show that that kz − yk ≥ θ for every y ∈ Y . We have
kz − yk = kc(v − y0 ) − yk
= c v − y0 − c−1 y
= c kv − y1 k ,
Theorem 0.17.5. If a normed space X has the property that the closed unit ball M = {x : kxk ≤
1} is compact, then X is finite dimensional.
Proof. We assume that M is compact but dimX = ∞, and show that this leads to a contradiction.
We choose any x1 of norm 1, that is kx1 k = 1. This x1 generates a one dimensional subspace
X1 = {αx1 } of X, which is closed and is a proper subspace of X since dim X = ∞. By Riesz’s
lemma there is an x2 ∈ X of norm 1 such that kx2 − x1 k ≥ θ = 1/2. In particular, kx3 − x1 k ≥
1/2, kx3 − x2 k ≥ 1/2. Proceeding by induction, we obtain a sequence (xn ) of elements xn ∈
M such that kxm − xn k ≥ 1/2. Obviously, (xn ) can’t have a convergent subsequence. This
contradicts the compactness of M. Hence our assumption dimX = ∞ is false and dimX < ∞.
49
0.18. BANACH FIXED POINT THEOREM AND APPLICATIONS
for all x, y ∈ X.
Example 0.18.1. Let X = R be the space of reals with usual metric. Let T : X → X be defined
by
(a) T (x) = x/3, for all x ∈ X
(b) T (x) = 2x, for all x ∈ X
Then for x, y ∈ X, (a) is contraction but (b) is not.
(a)
1
|T (x) − T (y)| = |x/3 − y/3| = |x − y|.
3
So, T is a contraction mapping.
(b)
|T (x) − T (y)| = |2x − 2y| = 2|x − y|.
So, T is not a contraction mapping.
Proof. Let (X, d) be a metric space and let T : (X, d) → (X, d) be a contraction mapping. So
there exists a constant 0 < α < 1 such that
for all x, y ∈ X.
Let ε > 0 be arbitrary. Choose δ = αε (> 0), such that d(x, y) < δ . So
ε
d(T (x), T (y)) < α = ε.
α
Therefore, T is uniformly continuous on X.
Proof. Let (X, d) be a complete metric space and f : X → X is a contraction mapping. For any
x0 ∈ X, define a sequence (xn ) in X such that
xn = f (xn−1 ) = f n (x0 ) n ≥ 1.
50
0.18. BANACH FIXED POINT THEOREM AND APPLICATIONS
d(xn+1 , xn ) ≤ α n d(x1 , x0 ).
where ai j and bi are real coefficients such that ∑nj=1 |ai j | ≤ α < 1 for each i = 1, ..., n.
Take x = (x1 , x2 , ..., xn ) ∈ Rn and y = (y1 , y2 , ..., yn ) ∈ Rn . Define a metric d on Rn by
51
0.18. BANACH FIXED POINT THEOREM AND APPLICATIONS
Define T : Rn → Rn by T (x) = y.
(1) (1) (1) (2) (2) (2)
Let x(1) , x(2) ∈ Rn such that x(1) = (x1 , x2 , ..., xn ) and x(2) = (x1 , x2 , ..., xn ).
(1) (1) (1) (1) (1) (1)
Take T (x(1) ) = y(1) and T (x(2) ) = y(2) , where y(1) = (y1 , y2 , ..., yn ) and y(2) = (y1 , y2 , ..., yn ).
(1) (1) (2) (2)
Here yi = ∑nj=1 ai j x j + bi , and yi = ∑nj=1 ai j x j + bi , for i = 1, ..., n. Consider
(1) (2)
d(y(1) , y(2) ) = max |yi − yi |
i
n n
(1) (2)
= max | ∑ ai j x j − ∑ ai j x j |
i j=1 j=1
n
(1) (2)
= max | ∑ ai j (x j − x j )|
i j=1
n
(1) (2)
≤ max ∑ |ai j | |x j − x j |
i j=1
(1) (2)
≤ α max |x j − x j |
j
= αd(x(1) , x(2) ).
Hence d(T (x(1) ), T (x(2) )) ≤ αd(x(1) , x(2) ), that is T is a contraction mapping. Hence by Ba-
nach contraction mapping theorem T has a unique fixed x in Rn .
in the neighbourhood [x0 − t, x0 + t] such that y0 = φ (x0 ). Now consider a closed circular disc
U centred at (x0 , y0 ) with some positive radius such that U is a subset of D.
Now, f is continuous over U. Let m = sup | f (x, y)| < ∞.
(x,y)∈U
Now, choose two positive reals t and δ such that
(i) 0 < t < 1/L
(ii) mt ≤ δ
52
0.18. BANACH FIXED POINT THEOREM AND APPLICATIONS
53
0.18. BANACH FIXED POINT THEOREM AND APPLICATIONS
So,
Z x Z x
|ψ1 (x) − ψ2 (x)| = | f (t, φ1 (t))dt − f (t, φ2 (t))dt|
x x0
Z 0x
=| ( f (t, φ1 (t)) − f (t, φ2 (t)))dt|
x
Z x0
≤ | f (t, φ1 (t)) − f (t, φ2 (t))|dt
x0
Z x
≤L |φ1 (t) − φ2 (t)|dt
x0
Z x
≤ L d(φ1 , φ2 ) dt
x0
= L|x − x0 |d(φ1 , φ2 )
≤ L t d(φ1 , φ2 ), 0 < Lt < 1.
d(ψ1 , ψ2 ) ≤ L t d(φ1 , φ2 )
or it can be rewritten as d(T (φ1 ), T (φ2 )) ≤ L t d(φ1 , φ2 ). So, T is a contraction map. Hence
by Banach Contraction Principle Theorem T has a unique fixed point in E say φ ∈ E. i.e.
T (φ ) = ψ. So φ (x) = y0 + xx0 f (t, φ (t))dt, where y0 = φ (x0 ).
R
54
0.19. INNER PRODUCT AND HILBERT SPACES
55
0.19. INNER PRODUCT AND HILBERT SPACES
of the inner product which are indeed immediate consequences of the axioms (I1)-(I4) of Def-
inition, and can be verified easily. If OV denotes the zero vector in V and 0 denotes the scalar
zero in F, then, for any vector v ∈ V , we have 0.v = OV so that by the linearity property of the
inner product h0V , vi = h0.v, vi = 0 hv, vi = 0.
Given a vector space V over F, a semi-inner product for V is a function on V × V which
satisfies all the axioms of the inner product except (I4) and instead of (I4), satisfies just the
condition hu, ui > O. Thus, a semi-inner product allows the possibility that for some u 6= 0,
hu, ui = O.
Proposition 0.19.1. In an inner product space (V, h., .i), we have for each u, v, w ∈ V and α, β ∈
F
(i) h0, ui = hu, 0i = 0
(ii) hαu + β v, wi = α hu, wi + β hv, wi [Linearity]
(iii) hu, αv + β wi = α hu, vi + β hu, wi [Conjugate Linearity]
The bar over α, β in (iii) and also the bar over βk in (0.19) are to be omitted in the case of a
real space as it has no effect.
From Definition and Proposition, we say that the key properties that characterize the inner
product are ”Hermitian symmetric, conjugate bilinear and positive definite”. Clearly, for ev-
ery w ∈ V , hu, wi = hv, wi holds iff hu − v, wi = 0, which for the choice w = u − v ∈ V gives
hu − v, u − vi = 0, that is u = v. The simplest example of an inner product space is C itself,
with hu, vi = uv.
Schwarz’s Inequality:
We may observe that the dot product a.b = a1 b1 + a2 b2 + a3 b3 , where a, b ∈ R3 is inner
product on R3 and the absolute value |a| = (a.a)1/2 , where a ∈ R3 , defines a norm on R3 .
Definition 0.19.2. If V is an inner product space equipped with an inner product h., .i, the
norm k.k on V associated with the inner product is the nonnegative real number defined by the
√
formula kuk = u, u for all u ∈ V . Here, it is evident that for every u ∈ V and α ∈ C, we have
the homogeneity condition kαuk2 = hαu, αui = αα hu, ui = |α|2 kuk2 .
Hence kuk ≥ 0 with equality iff u = 0. This property is called positivity condition of the
norm. p
For z = (z1 , z2 , ..., Zn ) ∈ Cn , kzk = hz, zi = ∑nk=1 |zk |2 which we have already noticed
p
as the Euclidean distance from the origin to the point z ∈ Cn . This observation shows that
the definition of inner product space arise naturally as a generalization from the concept of
(Euclidean) length that is familiar for tne set of real or complex numbers.
p naturally arises is: given an inner product h., .i on a vector space V is the
A question that
map x 7→ kxk := hx, xi, x ∈ V , a norm on V ??
To prove this we use Schwarz’s inequality
56
0.19. INNER PRODUCT AND HILBERT SPACES
Theorem 0.19.2. If h., .i is an inner product on a linear space V and x, y ∈ V , then |hx, yi| ≤
kxk kyk.
Proof. The above result is obvious if y = 0. Hence assume that y 6= 0. Then for every α ∈ F,
we have kx − αyk2 ≥ 0. Therefore,
kx − αyk2 = hx − αy, x − αyi
= kxk2 − hx, αyi − hαy, xi + |α|2 kyk2
= kxk2 − 2Re hx, αyi + |α|2 kyk2 .
By taking α = hx, yi / kyk2 , we have |α|2 kyk2 = |hx, yi|2 / kyk2 and hx, αyi = α hx, yi = |hx, yi|2 / kyk2 ,
so that 0 ≤ kx − αyk2 = kxk2 − |hx, yi|2 / kyk2 . Hence, |hx, yi| ≤ kxk kyk.
p p
If x and y are two vectors in an inner product space, then |hx, yi| = hx, xi hy, yi
p
Theorem 0.19.3. If h., .i is an inner product on a linear space X, then x 7→ kxk := hx, xi, x ∈ X
is a norm on X.
Proof. As h., .i is an inner product on a linear space X, using Schwarz inequality, we get
kx + yk2 = hx + y, x + yi
= hx, xi + hx, yi + hy, xi + hy, yi
= kxk2 + kyk2 + 2Re hx, yi
≤ kxk2 + kyk2 + 2|hx, yi|
≤ kxk2 + kyk2 + 2 kxk kyk
= (kxk + kyk)2 .
Thus k.k satisfies the inequality kx + yk ≤ kxk + kyk for every x, y ∈ X. It is easy to verify the
other conditions.
Hence inner product spaces are normed spaces, and Hilbert spaces are Banach spaces. But not
all normed spaces are inner product spaces.
Theorem 0.19.4. If x and y are two vectors in inner product space, then
(i) kx + yk2 + kx − yk2 = 2 kxk2 + 2 kyk2 [Parallelogram law]
(ii) 4 hx, yi = kx + yk2 − kx − yk2 + i kx + iyk2 − i kx − iyk2 .
[Polarisation Identity]
Proof. (i) x and y are two vectors in inner product space
kx + yk2 = hx + y, x + yi
= hx, xi + hx, yi + hy, xi + hy, yi
= kxk2 + kyk2 + hx, yi + hy, xi .
Also, we have
kx − yk2 = hx − y, x − yi
= hx, xi − hx, yi − hy, xi + hy, yi
= kxk2 + kyk2 − hx, yi − hy, xi .
57
0.19. INNER PRODUCT AND HILBERT SPACES
Theorem 0.19.5. Hilbert space or inner product is jointly continuous, that is, xn → x, yn → y,
=⇒ hxn , yn i → hx, yi.
Proof. Consider
Now, xn → x and yn → y, we have kxn − xk → 0 and kyn − yk → 0. Hence |hxn , yn i−hx, yi| → 0,
or hxn , yn i → hx, yi.
Definition 0.19.3. A complete inner product space is called a Hilbert space i.e. an inner product
space X which is complete with respect to a metric d : X × X → R induced by the inner product
1
h, i on X × X i.e. d(x, y) = hx − y, x − yi 2 for all x, y ∈ X.
Theorem 0.19.6. A Banach space X is a Hilbert space if and only if parallelogram law holds
in it.
Proof. We know that every Hilbert space X is a Banach space where parallelogram law holds
in it.
Conversely suppose that X is a Banach space where parallelogram law holds. Without
loss of generality we can assume a function h, i whose range is R. For all x, y ∈ X, define
h, i : X × X → R by
1h 2 2
i
hx, yi = kx + yk − kx − yk .
4
Here hx, yi = hx, yi as hx, yi is real. Also hx, xi ≥ 0 and hx, xi = 0 iff x = 0.
Now, for u, v, w ∈ X, we have
and
ku − v + wk2 + ku − v − wk2 = 2(ku − vk2 + kwk2 ).
On subtracting these two, we get
58
0.19. INNER PRODUCT AND HILBERT SPACES
hnx, yi = h−mx, yi
= m h−x, yi
= −m hx, yi
n hx, yi .
Hence hrn x, yi → hλ x, yi. So, hlambdax, yi = λ hx, yi for all x, y ∈ X. So, X is an inner product
space with respect to h, i consequently, X is a Hilbert space.
hx, yi = ξ1 η1 + ... + ξn ηn ,
59
0.19. INNER PRODUCT AND HILBERT SPACES
p q
where x = (ξ1 , ..., ξn ) and y = (η1 , ..., ηn ). Also, kxk = hx, xi = ξ12 + ... + ξn2 and Eu-
p p
clidean metric defined by d(x, y) = kx − yk = hx − y, x − yi = (ξ1 − η1 )2 + ... + (ξn − ηn )2 .
With respect to this metric we can at once see that Rn is complete so as to make Rn , a Hilbert
space.
Proof. We define the inner product by hx, yi = ∑∞j=1 ξ j η j . Here, we consider kxk = hx, xi1/2 =
q
∑∞j=1 |ξ j |2 . Here it is easily shown that all inner product axioms are satisfied in `2 . The
p 1
metric d of `2 is defined by d(x, y) = kx − yk = hx − y, x − yi = ∑|ξi − η j | 2 is complete so
as to make `2 a Hilbert space.
Proposition 0.19.9. For 1 ≤ p < ∞, ` p with p 6= 2 is not an inner product space, hence not
Hilbert space.
Proof. A norm on an inner product space satisfies the parallelogram equality kx + yk2 +kx − yk2 =
2 kxk2 + 2 kyk2 . Here ` p with p 6= 2 does not satisfies this, for example, x = (1, 1, 0, 0, ...) ∈ ` p
and y = (1, −1, 0, ...) ∈ ` p . Then kxk = kyk = 21/p and kx + yk = kx − yk = 2. Hence parallel-
ogram equality fails if p 6= 2.
But ` p is complete. Hence ` p with p 6= 2 is a Banach space which is not Hilbert space.
Proposition 0.19.10. C[a, b] is not inner product and hence not Hilbert space.
Proof. Define kxk = supt∈[a,b] |x(t)| does not satisfy the parallelogram equality kx + yk2 +
kx − yk2 = 2 kxk2 + 2 kyk2 . Take x(t) = 1 and y(t) = (t − a)/(b − a), we have kx(t)k = 1
and ky(t)k = 1. x(t) + y(t) = 1 + (t − a)/(b − a) and x(t) − y(t) = 1 − (t − a)/(b − a). Hence
kx + yk = 2, kx − yk = 1 and kx + yk2 + kx − yk2 = 5. But 2 kxk2 + 2 kyk2 = 4. Hence the
result.
60
0.19. INNER PRODUCT AND HILBERT SPACES
Definition 0.19.4. Isomorphism. An isomorphism T if an IPS (X1 , hi1 ) onto an IPS (X2 , hi2 )
over the same field K is a bijective linear operator T : X1 → X2 which preserves the inner
product, that is, hT x, Tyi = hx, yi.
Definition 0.19.5. A subspace Y of an inner product space X is defined to be a vector subspace
of X taken with the inner product on X restricted to Y ×Y .
Similarly, a subspace Y of a Hilbert space H is defined to be a subspace of H, regarded as an
inner product space. Note that Y need not be a Hilbert space because Y may not be complete.
Note. Let Y be a subspace of a Hilbert space H. Then:
(a) Y is complete if and only if Y is closed in H.
(b) If Y is finite dimensional, then Y is complete.
(c) If H is separable, so is Y . More generally, every subset of a separable inner product
space is separable. [It has countable subset which is dense in itself]
Definition 0.19.6. An element x of an inner product space X is said to be orthogonal to an
element y ∈ X if hx, yi = O. We also say that x and y are orthogonal, and we write x ⊥ y.
Law of cosines in R2 is ku − vk2 = kuk2 + kvk2 − 2 kuk kvk cos θ , where u, v ∈ R2 and θ is
angle between u and v. hu − v, u − vi = hu, ui + hv, vi − 2 kuk kvk cos θ . On simplifying, we get
−2 hu, vi = −2 kuk kvk cos θ or cos θ = hu, vi /(kuk kvk).
If u ⊥ v, then cos θ = 0, implies that hu, vi = 0. Conversely, if hu, vi = 0, we have cos θ = 0,
θ = π/2, u ⊥ v.
Example 0.19.11. (1) The space Rn is a Hilbert space with inner product definedq by hx, yi =
p
ξ1 η1 +...+ξn ηn , where x = (ξ1 , ..., ξn ) and y = (η1 , ..., ηn ). Also, kxk = hx, xi = ξ12 + ... + ξn2
p p
and Euclidean metric defined by d(x, y) = kx − yk = hx − y, x − yi = (ξ1 − η1 )2 + ... + (ξn − ηn )2 .
It is complete.
In trigonometry, using law of cosines, we have ka − bk2 = kak2 + kbk2 − 2 kak kbk cos θ ,
where θ is acute angle between a and b
”Is it possible to generalize the concept of angle between vectors in an inner product space
”
ku − vk2 = kuk2 + kvk2 − 2Re hu, vi. By Schwarz inequality, |Re hu, vi| ≤ |hu, vi| ≤ kuk kvk,
so that −1 ≤ Rehu,vi Rehu,vi 2 2
kukkvk ≤ 1. Therefore, if we define cos φ = kukkvk , then ku − vk = kuk + kvk −
2
61
0.20. ORTHOGONAL AND ORTHONORMAL VECTORS
Proof. (i) Let x ∈ M2⊥ . So hx, yi = 0, for all y ∈ M2 . So hy, xi = 0 for all y ∈ M2 . Therefore,
hy, xi = 0, for all y ∈ M1 as M1 ⊂ M2 . So x ∈ M1⊥ . Hence M1⊥ ⊃ M2⊥ .
(ii) Let x ∈ M. Then hy, xi = 0 for all y ∈ M ⊥ . So x ∈ (M ⊥ )⊥ . Thus M ⊂ M ⊥⊥ .
(iii) By (ii) M ⊂ M ⊥⊥ . So by (i) (M ⊥⊥ )⊥ ⊂ M ⊥ . Again by (ii), M ⊥ ⊂ (M ⊥ )⊥⊥ ⊂ (M ⊥⊥ )⊥ .
Hence (M ⊥⊥ )⊥ = M ⊥ .
(iv) Let x, y ∈ M ⊥ and α be any complex scalar. Therefore x ⊥ M, y ⊥ M, (x − y) ⊥ M and
αx ⊥ M and so x − y, αx ∈ M ⊥ . Consequently M ⊥ is a linear subspace of H.
62
0.20. ORTHOGONAL AND ORTHONORMAL VECTORS
Let x ∈ M ⊥ . So there exists a sequence (xk ) in M such that limk xk = x with xk ⊥ M. By the
continuity of inner product we see that x ⊥ M. Hence M ⊥ is closed.
Theorem 0.20.4. (Dense set). For any subset M 6= 0/ of a Hilbert space H, the span of M is
dense in H if and only if M ⊥ = {0}.
An orthogonal sequence in X is (un ), where un (t) = cos nt, n = 0, 1, ... Another orthogonal
sequence in X is (vn ) , where vn (t) = sin nt,
n = 1, 2, ....
R 2π 0, m 6= n
Now hum , un i = 0 cos mt cos nt dt = π, m = n = 1, 2, ...
2π, m = 0 = n
√
Hence an orthonormal sequence is (en ), where e0 (t) = 1/ 2π, en (t) = ukun (t) nk
= cos√ nt , n =
π
1, 2, ...
Similarly for vn , we can obtain an orthonormal sequence eˆn as eˆn = vkvn (t)
nk
= sin
√ nt , n = 1, 2, ...
π
63
0.20. ORTHOGONAL AND ORTHONORMAL VECTORS
Proof. Let {en }, n = 1, 2, . . . be an orthonormal system in H and let for a finite subset {e1 , e2 , . . . , en }
of the system we have α1 e1 +α2 e2 +· · ·+αn en = θ , where αi ’s are scalars. Then for 1 ≤ j ≤ n,
0 = θ,ej = ∑ αiei, e j = ∑ αi ei , e j = α j e j , e j = α j ,
Theorem 0.20.7. If M and N are two closed subspace of a Hilbert space H with M ⊥ N then
M + N is closed in H.
Proof. As H is complete and M and N are closed so M and N are complete. Since M ⊥ N, so
M ∩ N = {θ }. So H = M N.
L
So ||zn − zm ||2 = ||xn − xm ||2 + ||yn − ym ||2 . Since zk → z, {zk } is a Cauchy sequence in H. So
||zn −zm || → 0 as n, m → ∞. So ||xn −xm ||2 +||yn −ym ||2 → 0. Hence {xn } and {yn } are Cauchy
sequences in M and N respectively and xn → x ∈ M and yn → y ∈ N (by completeness of M
and N). Since zn = (xn + yn ) → (x + y) ∈ M + N as and so z = x + y ∈ M + N. Consequently
M + N is closed in H.
64
0.21. BEST APPROXIMATION IN HILBERT SPACE
δ = inf kx − yk = kx − yk .
y∈M
ky − y0 k2 = ky − x − (y0 − x)k2
= 2 ky − xk2 + 2 ky0 − xk2 − k(y − x) + (y0 − x)k2
2
1
= 2δ 2 + 2δ 2 − 22 (y + y0 ) − x .
2
65
0.21. BEST APPROXIMATION IN HILBERT SPACE
Turning from arbitrary convex sets to subspaces, we obtain a following result which gener-
alizes the familiar idea of elementary geometry that the unique point y in a given subspace Y
closest to a given x is found by ”dropping a perpendicular from x to Y .”
Projection Theorem.
Lemma 0.21.2. (Orthogonality). In Theorem 0.21.1, let M be a complete subspace Y and x ∈ X
fixed. Then z = x − y is orthogonal to Y , that is, z ⊥ Y .
Proof. Suppose that z = x − y is not orthogonal to Y , there would be a y1 ∈ Y such that
hz, y1 i = β 6= 0.
β
Further, if we choose β − α hy1 , y1 i = 0, that is α = hy1 ,y1 i , we have
β |β |2
kz − αy1 k2 = kzk2 − β − α[0] = δ 2 − < δ 2.
hy1 , y1 i hy1 , y1 i
But this is impossible because we have z − αy1 = x − y2 where y2 = y + αy1 ∈ Y so that
kz − αy1 k ≥ δ . Therefore, we must have z ⊥ Y .
66
0.21. BEST APPROXIMATION IN HILBERT SPACE
Our goal is a representation of a Hilbert space as a direct sum which is particularly simple
and suitable because it makes use of orthogonality. To understand the situation and the prob-
lem, let us first introduce the concept of a direct sum. This concept makes sense for any vector
space and is defined as follows.
Definition 0.21.1. (Direct sum). A vector space X is said to be the direct sum of two subspaces
Y and Z of X, written X = Y Z, if each x ∈ X has a unique representation x = y + z, y ∈ Y, z ∈
L
and for all x ∈ H there exists y ∈ Y such that x = y + z, where z ∈ Y ⊥ . Then the orthogonal
projection onto Y is P : H → Y given by x 7→ y = Px. P is called the (orthogonal) projection
(or projection operator) of H onto Y .
Clearly P is bounded. Further P is idempotent, that is, P2 = P, which we call idempotent.
Definition 0.21.4 (Annihilator). An orthogonal complement is a special annihilator, where, by
definition, the annihilator M ⊥ of a set M 6= 0/ in an inner product space X is the set M ⊥ = {x ∈
X : x ⊥ M}.
Thus, x ∈ M ⊥ if and only if hx, vi = 0 for all v ∈ M.
Remark 0.21.1. M ⊥ is a vector space, since x, y ∈ M ⊥ implies for all v ∈ M and scalars α, β ,
we have hαx + β y, vi = α hx, vi + β hy, vi = 0, hence αx + β y ∈ M ⊥ .
M ⊥ is closed.
M ⊂ (M ⊥ )⊥ = M ⊥⊥ , because x ∈ M =⇒ x ⊥ M ⊥ =⇒ x ∈ M ⊥⊥ .
Lemma 0.21.5. (Dense set). For any subset M 6= 0/ of a Hilbert space H, the span of M is dense
in H if and only if M ⊥ = {0}.
67
0.21. BEST APPROXIMATION IN HILBERT SPACE
Lemma 0.21.6. (Pythagorean Relation, Linear Independence). Let M = {x1 , ..., xn } be an or-
thogonal set. Then
(i) k∑ni=1 xi k2 = ∑ni=1 kxi k2 .
(ii) M is linearly independent.
68
0.21. BEST APPROXIMATION IN HILBERT SPACE
An orthogonal sequence in X is (un ), where un (t) = cos nt, n = 0, 1, ... Another orthogonal
sequence in X is (vn ) , where vn (t) = sin nt,
n = 1, 2, ....
R 2π 0, m 6= n
Now hum , un i = 0 cos mt cos nt dt = π, m = n = 1, 2, ...
2π, m = 0 = n
√
Hence an orthonormal sequence is (en ), where e0 (t) = 1/ 2π, en (t) = ukun (t) nk
= cos√ nt , n =
π
1, 2, ...
Similarly for vn , we can obtain an orthonormal sequence eˆn as eˆn = vkvn (t)
nk
= sin
√ nt , n = 1, 2, ...
π
Note that we even have um ⊥ vn for all m, n. These sequences appear in Fourier series.
69
0.22. SOME RESULTS IN INNER PRODUCT SPACES
where x ∈ X, (α1 , α2 , . . . , αn ) ∈ Cn , attains its absolute minimum value at one and only one
point (α1 , α2 , . . . , αn ) ∈ Cn , where hx, xk i = αk , k = 1, 2, . . . , n.
(ii) | hx, x1 i |2 + | hx, x2 i |2 + · · · + | hx, xn i |2 ≤ ||x||2 , x ∈ X.
Proof. (i) Let X be an inner product space and (ek ) be an orthonormal sequence in X. Consider
Further,
[ f (α1 , α2 , . . . , αn )]2
= ||x − ∑ αi xi ||2
= x − ∑ αi xi , x − ∑ αi xi
= hx, xi − x, ∑ αi xi − ∑ αixi, x + ∑ αixi, ∑ α j x j
= kxk2 − ∑ αi hx, xi i − ∑ αi hxi , xi + ∑ ∑ αi α j xi , x j
i j
70
0.22. SOME RESULTS IN INNER PRODUCT SPACES
Theorem 0.22.2. (Bessel’s Inequality). Let X be an inner product space and let (ek ) be an
orthonormal sequence in X. Let x ∈ H then we have
∞
∑ |hx, ek i|2 ≤ kxk2 .
k=1
Proof. Let X be an inner product space and let (ek ) be an orthonormal sequence in X, and
suppose x ∈ span({e1 , ..., en }) where n is fixed. Then we can represent x = ∑nk=1 αk ek as
* +
n
hx, ei i = ∑ αk ek , ei = αi
k=1
n
=⇒ x = ∑ hx, ek i ek .
k=1
Now let x ∈ X be arbitrary and take y ∈ span({e1 , ..., en }), where y = ∑nk=1 hx, ek i ek .
Define z by setting x = y + z =⇒ z ⊥ y because
hz, yi = hx − y, yi
= hx, yi − hy, yi
* +
n
= x, ∑ hx, ek i ek − kyk2
k=1
= ∑ hx, hx, ek i ek i − kyk2
= ∑ hx, ek i hx, ek i − ∑|hx, ek i|2 = 0.
Then, we have kxk2 = kyk2 + kzk2 or kzk2 = kxk2 − ∑|hx, ek i|2 ≥ 0. Therefore, ∑|hx, ek i|2 ≤
kxk2 . Hence the result.
Definition 0.22.1. (Fourier Coefficients). The sequence (hx, ek i) is called the Fourier coeffi-
cients of x with respect to (ek ).
Proposition 0.22.3. (Minimum property of Fourier Coefficients). Let {e1 , ..., en } be an or-
thonormal set in an in an inner product space X (n is fixed). Let x ∈ X be an arbitrary, fixed
element and let y = ∑nk=1 βk ek . Then kx − yk depends on β1 , ..., βn . Show by direct calculation
that kx − yk is minimized iff βi = hx, ei i, for all i = 1, ..., n.
71
0.22. SOME RESULTS IN INNER PRODUCT SPACES
These coefficients are called the Fourier coefficients of x. Obviously, the cosine and sine
functions in Fourier series are those of the sequences (uk ) and (vk ), where uk (t) = cos kt and
vk (t) = sin kt. Hence, we write the Fourier series as
∞
x(t) = a0 u0 (t) + ∑ (ak uk (t) + bk vk (t)).
k=1
We assume that termwise integration is permissible (uniform convergence would suffice) and
use the orthogonality of (uk ) and (vk ) as well as the fact that u j ⊥ vk for all j, k. Then we obtain
x, u j = a0 u0 , u j + ∑[ak uk , u j + bk vk , u j ]
(
2 2πa0 , j = 0
= a j u j, u j = a j u j =
πa j , j = 1, 2, ...
and
2
x, v j = b j v j = πb j , j = 1, 2, ...
−1
Solving for a j and b j , and using the orthonormal sequences (e j ) and (e j ), where e j = u j uj
−1
and e0j = v j v j , we obtain
1 1
aj = 2
x, u j = x, e j
uj uj
1 1
bj = 2
x, v j = x, e0j .
vj vj
Hence
∞
x(t) = hx, e0 i e0 + ∑ (hx, ek i ek + x, e0k e0k ).
k=1
For any orthonormal sequence (ek ) in a Hilbert space H, we may consider series of the
form
∑ αiei
where αi ’s are any scalars. Such a series converges and has the sum s if there exists an s ∈ H
such that the sequence (sn ) of the partial sums sn = ∑ni=1 αi ei converges to s, that is, ksn − sk →
0.
Theorem 0.22.4. Let (ek ) be a orthonormal sequence in a Hilbert space H. Then
(i) The series ∑ αi ei converges if and only if series ∑∞ 2
i |αi | converges.
(ii) If series ∑ αi ei converges, then the coefficients αi are the Fourier coefficients hx, ei i,
where x = ∑ αi ei . Hence in this case x = ∑ αi ei can be written as x = ∑ hx, ei i ei
(iii) for any x ∈ H the series x = ∑ αi ei with αi = hx, ei i converges.
Proof. (i) Let sn = ∑ni=1 αi ei and pn = ∑ni=1 |αi |2 . Then because of the orthonormality, for any
m and n > m,
ksn − sm k2 = kαm+1 em+1 + ... + αn en k2
= |αm+1 |2 + ... + |αn |2
= pn − pm .
72
0.22. SOME RESULTS IN INNER PRODUCT SPACES
Hence {sn } is Cauchy sequence in H if and only if (pn ) is Cauchy in R. Since H and R are
complete, we get the result.
(ii) Now take for j = 1, ...k and k ≤ n, we get sn , e j = α j . By assumption sn → x. Since
the inner product is continuous,
α j = sn , e j → x, e j .
Here we can take k(≤ n) as large as we please because n → ∞, so that we have α j = x, e j for
every j = 1, 2, ....
(iii) From the Bessel’s inequality, we see that the series ∑ | hx, ei i |2 converges. Therefore,
from (i), we get the result.
73
0.23. DUAL OF A HILBERT SPACE
Theorem 0.23.1. (Riesz Representation Theorem) Let H be a Hilbert space. Every bounded,
linear functional on H can be represented in terms of the inner product on H, i.e.,
kzk = k f k . (R2)
Theorem 0.23.2. Let H be a Hilbert space and H ∗ be its first conjugate space. Then for every
f ∈ H ∗ , there exists a unique z ∈ H such that
f (x) = hx, zi ,
74
0.23. DUAL OF A HILBERT SPACE
Example 0.23.3. Let H = L2 [0, 1] and D = C[0, 1] ⊂ L2 [0, 1]. Define linear functional T : D →
R as T f = f (0).
This operator is unbounded (and hence not-continuous), for take the sequence of functions,
(
1 − nx, 0 ≤ x ≤ 1n
fn (x) =
0, otherwise
kT fn k
Then k fn k → 0, where as kT fn k = | fn (0) = 1, that is, lim = ∞.
n→∞ k fn k
Example 0.23.4. Let H be a Hilbert space. Take y ∈ H and define the functional Ty = h., yi.
It is easy to see that it is linear. Also it is bounded as
| hy, yi | | hx, yi | kyk kxk
kyk = ≤ sup ≤ sup = kyk .
kyk x6=0 kxk x6=0 kxk
|Ty (x)|
Therefore, kT k = supx6=0 kxk = kyk .
Example 0.23.5. The Riesz representation does not hold for an incomplete inner product space.
Let X = c00 , the linear space of all scalar sequences each of which has only a finite number
of nonzero entries. For x, y ∈ X define hx, yi = ∑∞j=1 x( j)y( j).
p q
It is easy to see that h., .i is an inner product space and kxk2 = hx, xi = ∑∞j=1 |x( j)|2 , x ∈
X.
x(n)
Define f : X → K by f (x) = ∑∞ n=1 n , x ∈ X.
It is easy to see that f is linear. Also it is bounded by using Holder’s inequality, we have
∞
1 ∞
| f (x)|2 ≤ ( ∑ 2
)( ∑ |x(n)|2 ) ≤ C kxk2 .
n=1 n n=1
Hence f is continuous.
Also f ∈ X 0 has no representer in X.
To see this, let y ∈ X. As f (x) = hx, yi for all x ∈ X, then by taking en = (0, . . . , 0, 1, 0, . . .),
we found that y(n) = hen , yi = f (en ) = 1/n 6= 0 for all n ≥ 1. But this is impossible since
y ∈ c0 0. This shows that The Riesz representation does not hold for X.
Lemma 0.23.6. If hv1 , wi = hv2 , wi for all w in an inner product space X, then v1 = v2 . In
particular, hv1 , wi = 0 =⇒ v1 = 0 for all w ∈ X.
75
0.23. DUAL OF A HILBERT SPACE
Definition 0.23.1. Let X and Y be vector spaces over the same field K(= R or C). Then a
sesquilinear form (or sesquilinear functional) h on X ×Y is a mapping h : x ×Y → K such that
for all x, x1 , x2 ∈ X, y, y1 , y2 ∈ Y and all scalars α, β , we have
(a) h(x1 + x2 , y) = h(x1 , y) + h(x2 , y)
(b) h(x, y1 + y2 ) = h(x, y1 ) + h(x, y2 )
(c) h(α x, y) = αh(x, y)
(d) h(x, β y) = β h(x, y)
Hence h is linear in the first argument and conjugate linear in the second one.
Theorem 0.23.7. Let H1 , H2 be Hilbert spaces and h : H1 × H2 → K a bounded sesquilinear
form. Then h has a representation
Proof. We consider h(x, y) and this is linear in y because of bar. Keep x fixed. Then by Riesz
theorem on functionals, h(x, y) hy, zi. Hence
h(x, y) = hz, yi . (0.23.3)
here z ∈ H2 is unique but depends on fixed x ∈ H1 . It follows that (0.23.3) with variable x
defines an operator S : H1 → H2 given by Sx = z. So by substituting z = Sx in (0.23.3), we have
(0.23.1).
First, we prove S linear. In fact, its domain is the vector space H1 and from (0.23.1) and the
sesquilinearity we obtain
hS(αx1 + β x2 ), yi = h(αx1 + β x2 , y)
= h(αx1 , y) + h(β x2 , y)
= α hS(x1 ), yi + β hS(x2 ), yi
= hαS(x1 ) + β S(x2 ), yi ,
for all y ∈ H2 , so S(αx1 + β x2 ) = αS(x1 ) + β S(x2 ).
Now, we prove that S is bounded.
|h(x, y)|
khk = sup x, y 6= 0
kxk kyk
| hSx, yi |
= sup x, y 6= 0
kxk kyk
| hSx, Sxi |
≥ sup x, Sx 6= 0
kxk kSxk
||Sx||
= sup x 6= 0 = kSk .
kxk
This proves boundedness. Moreover ||h|| ≥ ||S||. We obtain (0.23.2) by noting ||h|| ≤ ||S||
follows by CS inequality
| hSx, yi | ||Sx||||y||
khk = sup x, y 6= 0 ≤ sup x, y 6= 0 = kSk .
kxk kyk kxk kyk
76
0.23. DUAL OF A HILBERT SPACE
77
0.24. ORTHOGONAL BASIS
Proof. Let S be a finite or infinite set of nonzero orthogonal vectors in the given space.
Let v1 , v2 , ..., vm be a set of m vectors in S, m ≤ dim S. Let v = α1 v1 + ... + αm vm , α ∈ F.
For any k, 1 ≤ k ≤ m, hv, vk i = ∑m i=1 αi hvi , vk i = αk hvk , vk i.
hv,vk i
Since vk 6= 0 hvk , vk i 6= 0. αk = hv ,v i = hv,vk2i , k = 1, 2, ..., m.
k k kvk k
If v = 0, then αk = 0 for k = 1, ..., m. Hence S is a LI set.
Corollary 0.24.3. If S = {v1 , ..., vn } is an orthogonal basis for V , then for v ∈ V the kth co-
ordinate of v is given by hv,vk2i , that is if v = (α1 , ...αn ) ∈ V , then αk hv,vk2i .
kvk k kvk k
Corollary 0.24.4. If V is an n-dimensional IPS then V can have atmost n mutually orthogonal
vectors.
78
0.24. ORTHOGONAL BASIS
n−1
vn = xn − ∑ hxn , ei i ei ,
i=1
vn
is not the zero vector and is orthogonal to e1 , e2 , ..., en . From it we obtain en = kvn k .
u.v
Definition 0.24.1. If u and v belong to inner product space X and v 6= 0, then vector v is
kvk2
called the vector projection of u along v.
Theorem 0.24.5. Let V be an IPS and {v1 , ..., vn } be a set of LI vectors in V . Then one can
constructa set {u1 , ..., un } of orthogonal vectors from {v1 , ..., vn } such that for any k, 1 ≤ k ≤ n,
{u1 , ..., uk } is a basis for L({v1 , ..., vn }).
On simplifying, we get
m vm+1 , u j
hum+1 , ur i = hvm+1 , ur i − ∑ 2
u j , ur
j=1 uj
= hvm+1 , ur i − hvm+1 , ur i = 0.
79
0.24. ORTHOGONAL BASIS
Example 0.24.7. Consider R3 with standard inner product. Let v1 = (3, 0, 4), v2 = (−1, 0, 7), v3 =
(2, 9, 11). {v1 , v2 , v3 } is LI set of vectors in R3 .
u1 = v1 = (3, 0, 4)
u2 = v2 − hv2 ,u12i u1 = (−1, 0, 7) − 25/25(3, 0, 4) = (−4, 0, 3)
ku1 k
hv3 ,u1 i
u3 = v3 − u − hv2 ,u22i u2 = (0, 9, 0)
ku1 k2 1 ku2 k
{u1 , u2 , u3 } = {(3, 0, 4), (−4, 0, 3), (0, 9, 0)} is an orthogonal set of vectors in R3 .
Further {ui / kui k : i = 1, 2, 3} = { 51 (3, 0, 4), 15 (−4, 0, 3), (0, 1, 0)} is an orthonormal set ob-
tained from {v1 , v2 , v3 }.
80
0.25. SOME EXAMPLES (IPS)
Example 0.25.3. Consider the functions sint and cost in the vector space C[−π, π] of contin-
uous functions on Rthe closed interval [−π, π]. Find hsint, costi .
π
hsint, costi = −π sint cost dt = 0.
Thus it is orthogonal.
Example 0.25.4. Find a nonzero vector w that is orthogonal to u1 = (1; 2; 1) and u2 = (2; 5; 4)
in R3 .
Let w = (x; y; z). Then we want hu1 , wi = 0 and hu2 , wi = 0. This yields the homogeneous
system
81
0.25. SOME EXAMPLES (IPS)
( (
x + 2y + z = 0 x + 2y + z = 0
2x + 5y + 4z = 0 y + 2z = 0
Here z is the only free variable in the echelon system. Set z = 1 to obtain y = −2 and x = 3.
Thus, w = (3, −2, 1) is a desired nonzero vector orthogonal to u1 and u2 .
Example 0.25.5. Let V be the vector space of polynomials P3 (t) with inner product h f , gi =
R1 2 3
−1 f (t)g(t) dt. Apply the Gram–Schmidt orthogonalization process to {1;t;t ;t } to find an
orthogonal basis { f0 ; f1 ; f2 ; f3 } with integer coefficients for P3 (t).
Solution. Now, we know that (
1 t r+s+1 1 2/(r + s + 1), r + s is even
ht r ,t s i = −1 t r+s dt = r+s+1
R
|−1 =
0, r + s is odd.
(1) First set f0 = 1.
ht,1i
(2) Compute t = h1,1i (1) = t − 0 = t. Set f 1 = t.
ht 2 ,1i ht 2 ,t i
(3) Compute t 2 − h1,1i (1) − ht,ti (t) = t 2 − 2/3 2 1
2 (1) + 0(t) = t − 3 .
Multiply by 3 to obtain f2 = 3t 2 − 1.
(4) Compute
3 t 3, 1 t 3 ,t t 3 , 3t 2 − 1
t − (1) − (t) − 2 (3t 2 − 1)
h1, 1i ht,ti h3t − 1, 3t 2 − 1i
2/5
= t 3 − 0(1) − (t) − 0(3t 2 − 1)
2/3
3
= t 3 − (t).
5
Multiply by 5 to obtain f3 (t) = 5t 3 − 3(t).
Thus {1,t, 3t 2 − 1, 5t 3 − 3t} is the required orthogonal basis.
Example 0.25.6. The vectors u1 = (1; 1; 0), u2 = (1; 2; 3), u3 = (1; 3; 5) form a basis S for Eu-
clidean space R3 . Find the matrix A that represents the inner product in R3 relative to this basis
S.
Solution. First compute ui , u j to obtain
82
0.26. TOTAL ORTHONORMAL SETS
83
0.26. TOTAL ORTHONORMAL SETS
{x, e1 , ...}. As x = ∑ ξi ei , we have hx, ei i = ξi , which implies ξi = 0 for all i. So x = (0, 0, 0, ...).
Hence A is complete.
Proposition 0.26.5. (Parseval’s equality) Show that an orthonormal set M in a Hilbert space H
is total iff
∑|hx, ek i|2 = kxk2 ,
k
for all x ∈ M.
Proof. If M is not total, there is a nonzero x ⊥ M in H. Since x ⊥ M we have 0 on the left-
hand side of Parseval’s Equality, whereas right hand side, which is equal to kxk2 6= 0. Hence if
Parseval’s equality holds for all x ∈ H, then M must be total in H.
Conversely, assume that M to be total in H. Consider any x ∈ H and its nonzero Fourier
coefficients arranged in a sequence, i.e., hx, e1 i , hx, e2 i .... Define
y = ∑ hx, ek i ek .
First we show that x − y ⊥ M. For every ei occuring in y = ∑ hx, ek i ek and using the orthonor-
mality, we have
84
0.27. ADJOINT OPERATOR
So T ∗ (y1 + y2 ) = T ∗ y1 + T ∗ y2 .
Now, we have to prove that T ∗ (αy) = α(T ∗ y) for y ∈ H and α ∈ C. Consider
hx, T ∗ (αy)i = hT x, αyi
= α hT x, yi
= α hx, T ∗ yi
= x, αT ∗ y .
85
0.27. ADJOINT OPERATOR
hT x, yi = hx, T ∗ yi .
Replacing T by T ∗ , we get
hT ∗ x, yi = hx, T ∗∗ yi .
Now interchanging x, y here we get hT ∗ y, xi = hy, T ∗∗ xi. Taking conjugates on both sides, we
have hx, T ∗ yi = hT ∗∗ x, yi . Hence hT x, yi = hx, T ∗ yi = hT ∗∗ x, yi , for all x, y ∈ H. Therefore
T = T ∗∗ .
Also, kT ∗ k ≤ kT k, replacing T by T ∗ we get kT ∗∗ k ≤ kT ∗ k. So, we have kT k ≤ kT ∗ k.
Therefore, kT k = kT ∗ k.
Note that the adjoint of identity operator I : H → H is the identity operator itself.
Similarly the adjoint of the Θ : H → H is the null operator itself.
Clearly I ∗ , Θ∗ ∈ BLd (H, H) with ||I|| = ||I ∗ || and ||Θ|| = ||Θ∗ ||.
Let H be a complex Hilbert space and let T1 and T2 be two bounded linear operators
mapping H onto itself i.e. T1 , T2 ∈ BLd (H, H). Then T1 +T2 and αT1 (α any complex scalar)
are also BLd (H, H).
Now
86
0.27. ADJOINT OPERATOR
Now
Theorem 0.27.3. For any T ∈ BLd (H, H), ||T ∗ T || = ||T ||2 = ||T T ∗ ||.
Proof. We always have ||T ∗ T || ≤ ||T ∗ ||||T || = ||T || ||T || = ||T ||2 , because T ∗ ∈ BLd (H, H)
and ||T || = ||T ∗ ||. Thus ||T ∗ T || ≤ ||T ||2 .
Again
Thus kT ∗ T k = kT k2 .
On replacing T by T ∗ and using T ∗ ∗ = T , we get kT T ∗ k = kT ∗∗ T ∗ k = k(T ∗ )∗ T ∗ k =
kT ∗ k2 = kT k2 . Hence the result.
87
0.28. SELF ADJOINT, NORMAL AND UNITARY OPERATORS
Theorem 0.28.2. Let T ∈ BLd (H, H) be such that for all x, y ∈ H, hT x, yi = 0. Then T equals
to the zero operator and conversely.
88
0.28. SELF ADJOINT, NORMAL AND UNITARY OPERATORS
Hence kT − T ∗ k and T ∗ = T .
Definition 0.28.2. Let H be a Hilbert space and T ∈ BLd (H, H) and T ∗ be adjoint of T . T is
said to be normal if T T ∗ = T ∗ T .
Remark 0.28.1. Every self adjoint operator is normal.
From the fact that T ∗∗ = T , we can say immediately that, if T is normal then its adjoint T ∗
is also normal.
Definition 0.28.3. A bounded linear operator T : H → H is said to be unitary if it satisfies the
condition T T ∗ = T ∗ T = I, where I is the identity operator on H.
If T is unitary, then T is injective. If T x1 = T x2 for x1 , x2 ∈ H then T ∗ T x1 = T ∗ T x2 and
since T ∗ T = I, we get x1 = x2 .
Also if T unitary then T is surjective. Because if y ∈ H then T (T ∗ y) = (T T ∗ )y = Iy = y.
Also it follows that unitary operators on H are precisely those operators whose inverses are
equal to their adjoints, that is T −1 = T ∗ .
Note that every self adjoint operator is normal. Also it is evident that every unitary
operator is normal.
A normal operator need not be self adjoint or unitary.
Example 0.28.6. Let I : H → H be an identity mapping and T = 2i I. Then T ∗ = (2iI)∗ = −2i I
and T −1 = −1 ∗ ∗ ∗ ∗
2 i I. So T T = T T = 4I. But T 6= T . Also T 6= T
−1 . So T is normal which is
Theorem 0.28.7. Let H be a Hilbert space, and T : H → H be unitary. Then T is isometry, that
is, hT x, T xi = hx, xi or kT xk = kxk, for all x ∈ H.
kT xk2 = hT x, T xi
= hx, T ∗ T xi
= hx, Ixi
= kxk2 .
Hence kT xk = kxk.
89
0.28. SELF ADJOINT, NORMAL AND UNITARY OPERATORS
Theorem 0.28.8. A bounded linear operator T on a Hilbert space H into itself is unitary if and
only if T is an isometrically isomorphism of H onto itself.
Proof. Suppose that T is an isometric and onto. Isometry implies injectivity, so T is bijective.
Hence T −1 exists and kT xk = kxk. Consider
hT ∗ T x, xi = hT x, T xi = hx, xi
h(T ∗ T − I)x, xi = 0.
T ∗ T = I.
90
0.29. LINEAR OPERATOR
Proof. (1) Let y1 , y2 ∈ R(T ) and α, β be arbitrary scalars. Since y1 , y2 ∈ R(T ), there exists
x1 , x2 ∈ D(T ) such that T x1 = y1 and T x2 = y2 . Since D(T ) is a vector space, αx1 + β x2 ∈
D(T ), so T (αx1 + β x2 ) ∈ R(T ). By the linearity of T , we have T (αx1 + β x2 ) = αT x1 +
β T x2 = αy1 + β y2 . Hence αy1 + β y2 ∈ R(T ), R(T) is a vector space.
(2) Choose n + 1 elements y1 , ..., yn of R(T ) such that yi = T xi for all i = 1, ..., n + 1 for
some x1 , ..., xn+1 ∈ D(T ).
Since dim D(T ) = n, so the set {x1 , ..., xn+1 } is linearly dependent. Hence α1 x1 + ... +
αn+1 xn+1 = 0, for some scalars α1 , ..., αn+1 not all zero. Since T is linear and T 0 = 0, we get
T (α1 x1 + ... + αn+1 xn+1 ) = α1 y1 + ... + αn+1 yn+1 = T (0) = 0.
This shows that {y1 , ..., yn+1 } is a linearly dependent set because the α’s are not all zero.Hence
dim R(T ) ≤ n.
(3) Take x1 , x2 ∈ N(T ). Then T x1 = 0 = T x2 . Since T is linear, for any scalars α, β we
have T (αx1 + β x2 ) = αT x1 + β T x2 = 0 + 0 = 0. This shows that αx1 + β x2 ∈ N(T ). Hence
N(T ) is a vector space.
into
Definition 0.29.3. A mapping T : D(T ) → Y is injective (or one-to-one) if for all x1 , x2 ∈ D(T )
such that x1 6= x2 =⇒ T x1 6= T x2 .
Equivalently, T x1 = T x2 =⇒ x1 = x2 .
So there exists the mapping T −1 : R(T ) → D(T ), y0 7→ x0 , that is y0 = T x0 which maps
every y0 ∈ R(T ) onto that x0 ∈ D(T ) for which T x0 = y0 . The mapping T −1 is called inverse
of T .
Therefore, we have T −1 T x = x for all x ∈ D(T ) and T T −1 y = y for all y ∈ R(T ).
91
0.29. LINEAR OPERATOR
Theorem 0.29.3. (Inverse operator) Let X,Y be vector spaces, both real or both complex. Let
T : D(T ) → Y be a linear operator with domain D(T ) ⊂ X and range R(T ) ⊂ Y . Then
(a) the inverse T −1 : R(T ) → D(T ) exists iff T x = 0 =⇒ x = 0.
(b) if T −1 exists, it is a linear operator.
(c) If dim D(T ) = n < ∞ and T −1 exists, then dim R(T ) = dim D(T ).
92
0.30. BOUNDED LINEAR TRANSFORMATION-I
Theorem 0.30.2. Let B(N, N1 ) be the set of all bounded (or continuous) linear transformations
of normed linear space N into normed linear space N1 . Then B(N, N1 ) is a normed linear space
with respect to pointwise linear operations (T + U)(x) = T (x) + U(x) and (αT )(x) = αT (x)
and norm defined by kT k = sup{kT (x)k : x ∈ N, kxk ≤ 1}. Further, if N1 is a Banach space,
then B(N, N1 ) is also a Banach space.
93
0.30. BOUNDED LINEAR TRANSFORMATION-I
Proof. We know that the set S of all linear transformations from a linear space to linear space
is also a linear space with pointwise linear operations. So to prove that B(N, N1 ) is linear it is
sufficient to prove that that B(N, N1 ) is a subspace of S.
Let T,U ∈ B(N, N1 ). Then T and U are bounded and so there exist real number k and l
such that kT (x)k ≤ k kxk and kU(x)k ≤ l kxk for all x ∈ N.
If α, β are scalars, then
k(αT + βU)(x)k = k(αT )(x) + (βU)(x)k
= kαT (x) + βU(x)k
≤ kαT (x)k + kβU(x)k
= |α| kT (x)k + |β | kU(x)k
≤ |α|k k(x)k + |β |l k(x)k
= [|α|k + |β |l] k(x)k .
Thus αT + βU ∈ B(N, N1 ). Thus B(N, N1 ) is a linear subspace of S.
Now, to prove that B(N, N1 ) is a normed linear space.
[n1] Since kT k = sup{kT (x)k : x ∈ N, kxk ≤ 1} and kT (x)k ≥ 0, thus kT k ≥ 0.
[n2] kT k = sup{kT (x)k : x ∈ N, kxk ≤ 1} = sup{kT (x)k / kxk : x ∈ N, x 6= 0}.
kT k = 0, =⇒ sup{kT (x)k / kxk : x ∈ N, x 6= 0} = 0 ⇔ kT (x)k = 0 or T = 0.
[n3] If α is any scalar, then kαT k = sup{k(αT )(x)k : x ∈ N, kxk ≤ 1} = sup{|α| kT (x)k :
x ∈ N, kxk ≤ 1} = |α| kT k.
[n4] If T,U ∈ B(N.N1 ), then
kT +Uk = sup{kT (x) +U(x)k : x ∈ N, kxk ≤ 1}
≤ sup{kT (x)k : x ∈ N, kxk ≤ 1} + sup{kU(x)k : x ∈ N, kxk ≤ 1}
= kT k + kUk
Hence B(N, N1 ) is normed linear space.
Now, to prove that B(N, N1 ) is complete if N1 is complete. Suppose that N1 is complete
and (Tn ) be a Cauchy sequence in B(N, N1 ). Then kTm − Tn k → 0. For each x ∈ N, we have
kTm (x) − Tn (x)k = k(Tm − Tn )(x)k ≤ kTm − Tn k kxk → 0. Hence (Tn ) is a Cauchy sequence in
N1 for each x ∈ N. Since N1 is complete so there exists a vector T (x) ∈ N1 such that Tn (x) →
T (x).
First we show that T is linear and bounded.
T (αx + β y) = lim Tn (αx + β y) = lim[αTn (x) + β Tn (y)] = αT (x) + β T (y).
Consider
kT (x)k = klim Tn (x)k = lim kTn (x)k ≤ lim kTn k kxk
≤ sup{kTn k kxk}
= sup{kTn k} kxk .
Also, we know that | kTm k − kTn k | ≤ kTm − Tn k → 0. Therefore (kTn k) is a Cauchy se-
quence of real numbers and hence convergent and bounded. So there exists k ≥ 0 such that
sup{kTn k} ≤ k. Hence kT (x)k ≤ k kxk. Thus T is bounded. Therefore T ∈ B(N, N1 ).
Finally, we have to show that Tn → T . Consider
kTn (x) − T (x)k = kTn (x) − Tm (x) + Tm (x) − T (x)k
≤ kTn (x) − Tm (x)k + kTm (x) − T (x)k
< ε/2 + ε/2 = ε.
94
0.30. BOUNDED LINEAR TRANSFORMATION-I
Theorem 0.30.3. Let T be a linear transformation from a normed linear space N to normed
linear space N1 . Then T is continuous at everypoint of N.
Proof. Let x1 and x2 be any two points of N. Suppose T is continous at x1 . Then for each
ε > 0, there exists δ > 0 such that kx − x1 k < δ , we have kT (x) − T (x1 )k < ε.
Now, for kx − x2 k < δ , k(x + x1 − x2 ) − x1 k < δ we have
kT (x + x1 − x2 ) − T (x1 )k < ε.
that is kT (x) − T (x2 )k < ε. This shows that T is continuous at x2 as well. Since x1 and x2
are arbitrary points, we have shown that if T is continuous at a particular point, then it is
continuous at all points.
95
0.31. BOUNDED AND CONTINUOUS LINEAR OPERATORS-2
kT k is called the norm of the operator or induced operator norm. If D(T ) = {0}, we define
kT k = 0; in this (relatively uninteresting) case, T = 0 since T 0 = 0.
Note that in the preceding definition (with c = kT k) implies that kT xk ≤ kT k kxk.
Note: kT k = supx∈D(T ),x6=0 (kT xk / kxk) minimum value of c, Norm of bounded linear
operator.
If T = 0, then kT k = 0.
Integral operator T : C[0, 1] → C[0, 1], x 7→ y = T x where y(t) = 01 k(t, τ)x(τ) dτ, where
R
kT k = sup kT xk.
x∈D(T ),x=1
Proof. (1) Write kxk = a and set y = (l/a)x, where x 6= O. Then kyk = kxk /a = 1, and since
T is linear, we have
kT xk x
kT k = sup = sup T( ) = sup kTyk.
x∈D(T ),x6=0 a x∈D(T ),x6=0 a y∈D(T ),y=1
96
0.31. BOUNDED AND CONTINUOUS LINEAR OPERATORS-2
Finally,
sup k(T1 + T2 )xk = sup k(T1 x + T2 x)k ≤ sup kT1 xk + sup kT2 xk .
kxk=1 kxk=1 kxk=1 kxk=1
Here k is a given function, which is called the kernel of T and is assumed to be continuous
on the closed square G = I × I in the tτ -plane, where I = [0, 1]. The operator is linear.
The continuity of k on the closed square implies that k is bounded, say, |k(t, τ)| ≤ k0 for
all (t, τ)(r, T ) ∈ G, where k0 is a real number. Furthermore, |x(t)| ≤ max|x(t)| = kxk .
Hence
kT xk = kyk
Z 1
= max| k(t, τ)x(τ) dτ|
t∈I 0
Z 1
≤ max| |k(t, τ)||x(τ)| dτ|
t∈I 0
≤ k0 kxk .
Hence T is bounded.
Theorem 0.31.3. If a normed space X is finite dimensional, then every linear operator on X is
bounded.
Proof. Let dim X = n and {e1 , ..., en } a basis for X. We take any x = ∑ni=1 αi ei and consider
any linear operator T on X. Since T is linear,
n n
kT xk = ∑ αiTei ≤ ∑|αi | kTei k ≤ max kTek k ∑|αi | = k ∑|αi |,
k 1 1
97
0.31. BOUNDED AND CONTINUOUS LINEAR OPERATORS-2
xi
Proposition 0.31.4. Define an operator `∞ → ∞ by T (xi ) = i for each (x) ∈ `∞ . Show that T
is linear and bounded.
Proof. Let (xi ), (yi ) ∈ `∞ . Then
αxi + β yi
T (αxi + β yi ) =
i
xi yi
= α +β
i i
= αT (xi ) + β T (yi ).
Hence T is linear.
Now, to show that T is bounded. Suppose that (x 6= 0) ∈ `∞ . Then
xi
kT (xi )k =
i
|xi |
= sup
i i
≤ sup|xi | = kxi k .
i
Hence T is bounded.
into
Lemma 0.31.5. Let T : D(T ) → Y be a bounded linear operator. Then
(a) (xn ) ⊂ D(T ), x ∈ D(T ) and xn → x then T xn → T x.
(b) The null space N(T ) is closed.
Theorem 0.31.6. Let T be a linear transformation from a normed linear space N to normed
linear space N1 . Then T is bounded iff it is continuous.
Proof. Let T be continuous and suppose that T is not bounded, that is, there exists no real
number k such that kT xk ≤ k kxk for every x ∈ N. Then for each positiveinteger n, we can find
kT xn k T xn xn
a vector xn such that kT xn k > n kxn k or nkxn k > 1 or nkx nk
> 1 or T nkx nk
> 1.
xn xn kxn k
Take yn = nkx nk
. kyn k = nkx nk
= nkx nk
= 1n → 0 and so yn → 0. Since kTyn k > 1, so
T (yn ) 6→ 0. Hence T is not continous at 0, which is contradiction. Hence T must be bounded.
98
0.31. BOUNDED AND CONTINUOUS LINEAR OPERATORS-2
Conversely, suppose that T is bounded such that there exists a real number k ≥ 0 such that
kT xk ≤ k kxk for all x ∈ N.
Now, we have to show that T is continuous. Let x ∈ N be arbitrary. For any ε > 0,
choose δ = ε/k. Then for all y ∈ N such that ky − xk < δ , we get kTy − T xk = kT (y − x)k ≤
k ky − xk < k(ε/k) = ε. Hence T is continuous at x. Since x is an arbitrary, so T is a continuous
mapping.
99
0.32. EXTENSION OF LINEAR OPERATOR
X Y
T
Theorem 0.32.1. (Bounded linear Extension) Let T : D(T ) → Y be a bounded linear operator,
where D(T ) ⊂ X, X a normed space and Y is a Banach space. Then T has an extension
T̂ : D(T ) → Y where T̂ is bounded linear operator of norm T̂ = kT k .
Proof. Let x ∈ D(T ). Then there exists a (xn ) ∈ D(T ) such that xn → x. Since T is linear and
bounded, we obtain
kT xn − T xm k = kT (xn − xm )k ≤ kT k kxn − xm k .
100
0.32. EXTENSION OF LINEAR OPERATOR
101
0.33. LINEAR FUNCTIONALS
Theorem 0.33.1. Let, X be a normed linear space over a field K, f : X → K be a linear func-
tional on X. Then f is continuous if and only if ker( f ) is closed in X.
102
0.33. LINEAR FUNCTIONALS
rx
|f( )| < 1
2||x||
r
| f (x)| < 1
2||x||
r
| f (x)| < 1
2||x||
2||x||
| f (x)| < .
r
So, f is bounded, implying that f is continuous.
Example 0.33.2. dot product f : Rn → R, f (x) = x.a, where
x = (x1 , x2 , x3 , . . . , xn ) ∈ Rn and a = (a1 , a2 , a3 , . . . , an ) ∈ Rn . Then f (x) is a bounded linear
functional and k f k = kak.
We have | f (x)| = |x.a| ≤ kxk kak.
k f k = sup{| f (x)| : x ∈ X kxk ≤ 1} ≤ kak
But k f k ≥ | fkak
(a)|
= |a.a|/ kak = kak2 / kak = kak. Hence k f k = kak.
functional, with k f k = b − a.
It is easy to see that f is linear.
Also k f k = sup| f (x)| = sup| ab x(t)dt| ≤ sup |x(t)|(b−a) ≤ kxk (b−a). Hence f is bounded
R
and k f k ≤ b − a.
Let x = x0 such that kx0 k = 1, | f (x0 )| = | ab dt| = (b − a). So k f k ≥ | fkx(x0k)| = b − a. Hence
R
0
k f k = b − a, length of the interval.
Example 0.33.4. X = `2 , p f (x) = ∑∞
ipξi αi is a bounded linear functional on `2 because | f (x)| =
∞ ∞ ∞ 2
| ∑i ξi αi | ≤ ∑i |ξi ||αi | ≤ ∑i |ξi | ∑∞ 2
i |αi | ≤ c kxk [By Cauchy Schwarz inequality].
Example 0.33.5. Unbounded linear functional equation: Take X = `∞ with norm defined by
kxk = supi |ξi |.
Here x ∈ (ξ1 , . . . , ξn , . . .)`∞ and ξi are scalars. Define f : X → R by f (x) = ∑ ξi . Then f is
unbounded linear functional.
Clearly f is linear. Choose the sequence {xn } ∈ `∞ where xn = ∑ni=1 ei , where ei = (0, 0, . . . , 1 , 0, . . .), i =
ithplace
1, 2, . . . . For each natural number n, note that ||xn || = || ∑ni=1 ei || = 1 and f (xn ) = n. Thus we
see that for each natural number n, | f (xn )| = n||xn ||. So it follows that || f || = ∞ and hence f is
an unbounded linear functional on `∞ .
The following result is required for the next theorem
Lemma 0.33.6. Let f be a non zero linear functional on a linear space X and let x0 ∈ X\ ker( f ).
Then any x ∈ X can be expressed uniquely in the form x = y + αx0 , where y ∈ ker( f ) and α is
scalar.
f (x) f (x)
Proof. Since x0 ∈ X\ ker( f ), f (x0 ) 6= 0. Choose α = f (x0 ) , x ∈ X and y = x − f (x0 ) x0 . Then
f (x) f (x)
x = y + αx0 . So f (y) = f (x − f (x0 ) x0 ) = f (x) − f (x0 ) f (x0 ) = 0. Hence y ∈ ker f .
103
0.33. LINEAR FUNCTIONALS
the result.
Theorem 0.33.7. Let X a normed linear space and f be a linear functional on X. Then the set
ker( f ) is either dense or closed in X.
Proof. Let us suppose that ker( f ) is not closed. Then there exists a point x0 ∈/ ker( f ) such that
x0 is a limit point of ker( f ). Since ker( f ) is a subspace of X, it follows that ker( f ) is also a
subspace of X. Therefore ker( f ) contains the linear span of x0 and ker( f ). Using previous
lemma, we get, X ⊂ ker( f ). Hence ker( f ) = X.
Using Theorems 0.33.1 and 0.33.7, we have the following result.
Corollary 0.33.8. Let, X be a normed linear space and f be a linear functional on X. Then f
is bounded (continuous) if and only if ker( f ) is not dense in X.
104
0.34. DUAL SPACE
g(α f1 + β f2 ) = gx (α f1 + β f2 )(x)
= (α f1 + β f2 )(x)
= α f1 (x) + β f2 (x)
= αgx ( f1 ) + β gx ( f2 ).
So gx is linear functional. So gx ∈ X ∗∗ .
To each x ∈ X there corresponds a gx ∈ X ∗∗ . So we define a mapping C : X → X ∗∗ by x 7→ gx
, x ∈ X Then C is called Canonical mapping of X into X ∗∗ . We claim C is also linear.
D(C) is a vector space and R(C) ⊂ X ∗∗ lies in Vector space and (C(αx + β y))( f ) =
gαx+β y ( f ) = f (αx + β y) = α f x + β f y = αgx ( f ) + β gy ( f ) = α(Cx) f + β (Cy)( f ). So C is
linear.
C is also called Canonical embedding of X into X ∗∗ .
L J
Definition 0.34.4. Isomorphism: Let (X, +, .) and (Y, , ) be two vector spaces. An isom-
porphism
J L J
T : X → Y is a bijective mapping which preserves the operation T (α.x + β .y) =
α T x β Ty.
Similarly if (X, d) and (Y, d1 ) be two metric spaces then an isomorphic T : X → Y is a
bijective mapping which preserves the distances, that is d(T x, Ty) = d(x, y), x, y ∈ X.
Definition 0.34.5. If X is isomorphic with a subspace of a vector space Y then we say that X is
embeddable in Y .
Since C : X → X ∗∗ x 7→ gx is a canonical embedding of X into X ∗∗ .
Definition 0.34.6. If R(C) = X ∗∗ , then X is said to be algebraically reflexive.
105
0.34. DUAL SPACE
Finite dimensional vector spaces are simpler than infinite dimensional ones, and it is
natural to ask what simplification this entails with respect to linear operators and func-
tionals defined on such a space. Linear operators on finite dimensional vector spaces can
be represented in terms of matrices, as explained below. In this way, matrices become the
most important tools for studying linear operators in the finite dimensional case.
Let X and Y be finite dimensional vector space over the same field. Let T : X → Y be
a linear operator. dim X = n and dimY = r Basis E = {e1 , ..., en } for X and B = {b1 , ..., br }
basis for Y arranged in definite order which we keep fixed.
Let x ∈ X, y = T x ∈ Y and Tei ∈ Y for all i has a unique representation
x = ∑ ξi ei (0.34.1)
Since T is linear
y = T x = T (∑ ξi ei ) = ∑ ξi T (ei ) (0.34.2)
We can say that T is uniquely determined if the images yi = Tei of the n-basis vectors
e1 , ..., en are prescribed. Since y, Tei ∈ Y , so
r
y = ∑ η jb j
j
or
r
Tei = ∑ τ ji b j
j
.
Consider y = ∑rj=1 η j b j = ∑ni=1 ξi T (ei ) = ∑ ξi ∑ τ j b j = ∑i (∑ j τ ji ξi )b j .
This implies that
η j = ∑ τ ji ξi (0.34.3)
i
106
0.35. EXAMPLES OF DUAL SPACE
Proof. F is a linearly independent set since ∑nk=1 βk fk (x) = 0 for all x ∈ X, with x = e j gives
∑nk=1 βk fk (e j ) = ∑nk=1 βk δ jk = β j = 0, so that all β j are zero. Now, we show that every f ∈ X ∗
can represented as a linear combination of the elements of F in a unique way. We write
f (e j ) = α j . Thus f (x) = ∑nj=1 ξ j α j for every x ∈ X. On the other hand, we obtain f j (x) =
f j (∑ni=1 ξi ei ) = ξ j . Hence, we get f (x) = ∑nj=1 α j f j (x). Hence the unique representation of the
arbitrary linear functional f on X terms of the functionals f1 , ... fn is f = α1 f1 + ... + αn fn .
Lemma 0.34.2. (Zero vector) Let X be a finite dlmensional vector space. If x0 ∈ X has the
property that f (x0 ) = 0 for all f ∈ X ∗ , then x0 = O.
Proof. Let {e1 ; ...; en } be a basis for X and let x0 = ∑ni=1 ξ0i ei . Then f (x0 ) = ∑ni=1 ξ0i αi . By
assumption f (x0 ) = 0, for all f ∈ X ∗ . Hence ξ0i = 0 for each i. Hence the result.
a+b+c = 1
a+b−c = 0
a−b−c = 0
107
0.35. EXAMPLES OF DUAL SPACE
p
Taking the supremum over all x of norm 1, we get k f k ≤ ∑ni=1 |αi |2 . Thus k f k ≤ kCk , ehere
C = (α1 , . . . , αn ) ∈ Rn . Thus for f ∈ X 0 , we get C = (α1 , . . . , αn ) ∈ Rn .
Conversely, C = (α1 , . . . , αn ) ∈ Rn . Define f : X → Rn by f (x) = f (x) = ∑ni=1 ξi αi , x =
∑ ξi ei , αi = f (ei ). So f is linear.
Now,
| f (x)| = | ∑ ξi αi |
≤ ∑ |ξi αi |
q q
≤ ∑ |ξi |2 ∑ |αi |2 = ||x||||C||.
f (x) = ∑ αi2
k f (x)k
kfk ≥
kxk
s
n
= ∑ |αi|2
i=1
p
k f k = ∑ni=1 |αi |2 .
This proves that the norm of f is the Euclidean norm, and k f k = kCk, where C = (αi ) ∈ Rn .
Hence the mapping of Rn0 onto Rn defined by f 7→ C = (αi ), αi = f (ei ) is norm preserving,
linear and bijective, it is an isomorphism. Hence the result.
Example 0.35.2. The dual/conjugate space of ` p , (1 < p < ∞) is `q ; here, 1 < p < ∞ and q is
the conjugate of p, that is, l/p + l/q = 1.
Solution. A Schauder basis for ` p is (ek ), where ek = (δki ) has 1 in the kth place and
zeros otherwIse. Then every x = (ξ1 , . . .) ∈ ` p has a unique representation x = ∑∞
k=1 ξk ek and
1
∑ |ξi | p < ∞, kxk = (∑ |ξi | p ) p .
We consider any f ∈ `0p , where `0p is the dual space of ` p .
Since f is linear and bounded, f (x) = ∑∞ k=1 ξk γk , where γk = f (ek ) are uniquely determined
by f .
(n)
Let q be(the conjugate of p and consider xn = (ξk ) with
(n) |γk |q /γk , i f k ≤ n, γk 6= 0,
ξk =
0, otherwise.
(n)n q
Hence f (xn ) = ∑∞
k=1 ξk γk = ∑k=1 |γk | .
Also, we know (q − 1)p = q, we have,
f (xn ) ≤ k f k kxn k
(n)
= k f k (∑ |ξk | p )1/p
= k f k (∑|γk |(q−1)p )1/p
= k f k (∑|γk |q )1/p .
108
0.35. EXAMPLES OF DUAL SPACE
| f (x)| = |∑ ξk γk |
≤ (∑ |ξk | p )1/p (∑ |γk |q )1/q
= kxk (∑ |γk |q )1/q .
Finally, we have to show that norm of f is the norm on the space `∞ . We have | f (x)| =
|∑ ξk γk | ≤ supi |γi | ∑ |ξk | = supi |γi | kxk .
Taking the supremum over all x of norm 1, we get k f k ≤ supi |γi |.
Hence we get k f k = supi |γi |, which is the norm on `∞ . Hence this formula can be writ-
ten k f k = kck∞ , where c = (γ j ) ∈ `∞ . It shows that the bijective linear mapping of `01 , onto
`∞ defined by f 7→ c = (γ j ) is linear and bijective, and norm preserving, so that it is an an
isomorphism.
109
0.36. LINEAR FUNCTIONALS AND HAHN-BANACH THEOREM
110
0.36. LINEAR FUNCTIONALS AND HAHN-BANACH THEOREM
Proof. Let M be the set of all linearly independent subsets of X. Since X 6= {0} so there exists
x 6= 0 and {x} ∈ M, then M 6= {0}.
Define inclusion relation on M, that is A ≤ B if A ≤ B. So M is Partially Oordered set.
We claim every chain C ⊂ M has an upper bound because union of all subset s of C are
elements of C. By Zorn’s lemma, M has a maximal element say B. We claim B is a Hamel
basis for X. Suppose it is not true, that is span B ⊂ X so there is a z ∈ X − spanB such that
B ∪ {z} would be Linearly independent set containing B as proper subset, which contradicts
thatB is maximal element. So spanB = X. Therefore B is a Hamel basis for X.
Theorem 0.36.3. In every Hilbert space H 6= {0} there exists a total orthonormal set.
Proof. Let M be set of all orthonormal subsets of H. Since H 6= {0} so there exists x 6= 0 and
x
an orthonormal subset of H is {y}, where y = kxk , then M 6= {0}.
On M define set inclusion relation, this gives M a POS. Pick a chain (TOS) C ⊂ M. This
will have an upper bound because union of all subsets in C.
By Zorn’s lemma, M will have a maximal element say F. We shall prove F is total in
H. Suppose F is not total in H, so there exists a nonzero z ∈ H such that z ⊥ F. Hence
z
F1 = F ∪ {e}, where e = kzk , is a orthonormal set which contains F as a proper subset. SO F
can not be maximal element. Therefore a contradiction, and shows that F is total in H.
Remarks 0.36.1. Hahn Banach Theorem is an extension theorem for linear functional on vec-
tor spaces.
This guarantes the normed space dual.
Uniform Boundedness Theorem (Banach-Steinhaus Theorem): This theorem gives condi-
tion sufficient for sequence (kTn k) to be bounded where Tn : X → Y , where X is Banach space
Y is normed space and Tn is bounded linear operator.
It has lot of applications like Summability theory, Analysis of Fourier series, weak conver-
gence
Open Mapping Theorem: This theorem says that a bounded linear operator T : X → Y X,Y
are Banach space is an open mapping, that is T sends open set to open set. Hence T is bijective
and continuous. [T is continuous iff T −1 G is open, G is open]
Closed graph theorem: It gives a condition under which a closed linear operator is bounded.
The operator T : X → Y (X,Y is normed space) is closed linear operator if the graph of T
(G(T ) = D(T ) ⊂ X → Y ) is closed in the normed space X × X.
Theorem 0.36.4. Hahn Banach Theorem Let M be a linear subspace of a normed linear space
N and f be a functional defined on M. Then f can be extended to a functional F defined on
the whole space N such that kFk = k f k.
HBT is discovered by Hahn in 1927 and in modified form by S. Banach in 1929 and
later extended to complex vector space by H. F. Bhonenhust et al in 1938. HBT is an
extension of linear functional. In HBT, the object to be extended is a linear functional f
which define on subspace Z of a vector space X and has a certain boundedness property
in terms of sublinear functional.
Definition 0.36.4. A Sublinear functional p(x) is a real valued functional on a vector space X
which is subadditive, that is, p(x + y) ≤ p(x) + p(y) for all x, y ∈ X and positive homogeneous,
that is, p(αx) = α p(x) for all α ≥ 0 in R and x ∈ X.
111
0.36. LINEAR FUNCTIONALS AND HAHN-BANACH THEOREM
Theorem 0.36.5. Hahn-Banach Theorem (real vector space): Let X be a real vector space and
p a sublinear functional on X. Further, let f be a linear functional defined on subspace Z of X
and satisfies f (x) ≤ p(x), for all x ∈ Z.
Then f has a linear extension fˆ : Z → X satisfying fˆ(x) ≤ p(x) for all x ∈ X, that is fˆ is a
linear functional on X, satisfies fˆ(x) ≤ p(x) on X and fˆ(x) = f (x) for any x ∈ Z.
Proof. Step 1. The set E of all linear extensions g of f satisfying the condition g(x) ≤ p(x) on
their domain D(g) can be partially ordered and Zorn’s lemma yields a maximal element fˆ of
E.
Step 2. fˆ is defined on the entire space X.
Step 3. An auxiliary relation which is used is step 2.
Proof of Step 1. Let E be the set of all linear extensions g of f which satisfies the condition
g(x) ≤ p(x) for all x ∈ D(g).
Clearly E 6= 0/ because f ∈ E ( f is linear and f (x) ≤ p(x), x ∈ D( f )). On E, we define a
partial order as follows g ≤ h if h is an extension of g, where g, h ∈ E, that is D(h) ⊃ D(g) and
h(x) = g(x) for all x ∈ D(g).
For any chain(TOS) C ⊂ E and define ĝ ∈ C by ĝ(x) = g(x) whenever x ∈ D(g).
Clearly ĝ is a linear functional [ĝ(αx + β y) = g(αx + β y) = αg(x) + β g(y), if αx + β y ∈
D(g) and g is linear] and domain D(ĝ) = ∪g∈C D(g). Obviously D(ĝ) is a vector space, as C is
a chain.
We claim that ĝ is an upper bound of C. Because, x ∈ D(g1 ) ∩ D(g2 ) with g1 , g2 ∈ C,
g1 (x) = g2 (x). Since C is chain so g1 ≤ g2 or g2 ≤ g1 . Hence g ≤ ĝ for all g ∈ C. So ĝ is an
upper bound of C.
Since C is a chain which has an upper bound and C is arbitrary chain of E. So by Zorn’s
lemma E has a maximal element fˆ, say. By the definition of E, fˆ is a linear extension of f
which satisfies fˆ(x) ≤ p(x), x ∈ D( fˆ).
Step 2. Now, we have to show that D( fˆ) = X. Suppose D( fˆ) 6= X, so we can choose a
y1 ∈ X − D( fˆ). Cleraly, y1 6= 0 as 0 ∈ D( fˆ).
Consider a subspace Y1 of X spanned by D( fˆ) and y1 . So any x ∈ Y1 can be written as
x = y + αy1 , y ∈ D( fˆ). We claim that this representation is unique, That is x = y + αy1 =
y + β y1 , y, y ∈ D( fˆ), so y − y = (β − α)y1 . But LHS y − y ∈ D( fˆ) and RHS y1 ∈/ D( fˆ). So only
possibility that y − y = 0 and β − α = 0, that is y = y.
Introduce a functional g1 on Y1 as follows g1 (x) = g1 (y + αy1 ) = fˆ(y) + αc, y ∈ D( fˆ) and
c is a real number. Then g1 is linear [y 7→ ay0 + by00 , fˆ is linear].
Further, for α = 0, g1 (y) = fˆ(y), so g1 is a proper extension of fˆ, that is an extension such
that D( fˆ) is a proper subset of D(g1 ).
If we prove that g1 (x) ≤ p(x) for all x ∈ D(g1 ), then g1 ∈ E and hence so that D( fˆ) = X
will contradict the maximality of fˆ, and our assumption D( fˆ) 6= X is wrong.
Step 3. Consider y, z ∈ D( fˆ). Then as fˆ is linear, we have
It implies that −p(−y1 −z)− fˆ(z) ≤ p(y+y1 )− fˆ(y), where y1 is fixed, y, z ∈ D( fˆ). Taking the
supremum of LHS and infimum of RHS, we get m0 = supz {−p(−y, −z) − fˆ(z)} ≤ infy {p(y +
y1 ) − fˆ(y)} = m1 , there exists c such that m0 ≤ c ≤ m1 , so −p(−y, −z) − fˆ(z) ≤ c z ∈ D( fˆ)
and c ≤ p(y + y1 ) − fˆ(y) for all y ∈ D( fˆ).
112
0.36. LINEAR FUNCTIONALS AND HAHN-BANACH THEOREM
Now, we have to prove that g1 (x) ≤ p(x) for all x ∈ D(g1 ), that is g(y + αy1 ) ≤ p(x). We
prove for cases when α > 0, α < 0 or α = 0.
Take α < 0. Replace z by α −1 y in −p(−y1 − z) − fˆ(z) ≤ c, we get −p(−y1 − α −1 y) −
ˆf (α −1 y) ≤ c. Multiply −α > 0, we get α[p(−y1 − α −1 y) + fˆ(α −1 y)] ≤ −αc, so −p(−αy1 −
y) ≥ fˆ(y) + αc Therefore p(x) = p(αy1 + y) ≥ g1 (x).
Take α = 0, result is obvious.
For α > 0, in a similar way, one can prove that g1 (x) ≤ fˆ(y) + αc ≤ p(x). Hence the result.
113
0.36. LINEAR FUNCTIONALS AND HAHN-BANACH THEOREM
Theorem 0.36.6. Generalized Form [Complex vector Space] Let X be a real or complex vector
space and p a real valued functional on X which is sub-additive, that is p(x + y) ≤ p(x) + p(y)
for all x, y ∈ X and for scalar α satisfies p(αx) = |α|p(x), x ∈ X.
Further, let f be a linear functional which is defined on a subspace Z of X and satisfying
| f (x)| ≤ p(x) for all x ∈ Z. Then f has a linear extension fˆ : Z → X satisfying | fˆ(x)| ≤ p(x)
for all x ∈ X.
Proof. (a) For real vector space: If X is real, then | f (x)| ≤ p(x) =⇒ f (x) ≤ p(x) for all x ∈ Z.
So by Hahn-Banach theorem (for real vector space), there is a linear extension fˆ from Z to X
such that f (x) ≤ p(x) for all x ∈ X.
Now to establish, | fˆ(x)| ≤ p(x). Consider − fˆ(x) ≤ fˆ(−x) ≤ p(−x) = | − 1|p(x) = p(x).
Thus | fˆ(x)| ≤ p(x)
(b) For complex vector space: Let X be a complex vector space. Then Z ⊂ X is also a
complex vector space. Hence f is complex valued linear functional defined on Z. So f (z) =
f1 (x) + i f2 (x), x ∈ Z and f1 , f2 are real valued.
Take X and Z as real vector spaces and denoted by Xr and Zr respectively, that is we restrict
multiplication by scalars to real numbers instead of complex numbers. Since f is linear on
Z, f1 , f2 are real valued and linear functionals defined on Zr . Further f1 (x) ≤ | f (x)|. So
f1 (x) ≤ p(x) for all x ∈ Zr . By HBT, there is a linear extension fˆ1 of f1 from Zr to Xr such that
for all x ∈ Xr .
Returning Z and X, and f = f1 + i f2 is linear on Z. For all x ∈ Z, i[ f1 (x)+ i f2 (x)] = i f (x) =
f1 (ix) + i f2 (ix). Equating real and imaginary parts, we have
So fˆ is linear.
Now, we’ll prove that fˆ satisfies | fˆ(x)| ≤ p(x) for all x ∈ X.
ˆ = 0 for any x then | fˆ(x)| ≤ p(x) as p(x) ≥ 0 [p(x + y) ≤ p(x) + p(y), p(αx) =
If f (x)
|α|p(x) =⇒ p(x) ≥ 0].
Let x be such that f (x)ˆ 6= 0. So using polar form f (x) ˆ = | f (x)|e
ˆ iθ =⇒ | f (x)| ˆ = f (x)e
ˆ −iθ =
f (xeˆ−iθ ). Since | f (x)|
ˆ is real so f (xeˆ−iθ ) is also real. Hence f (xeˆ−iθ ) = f1 (xeˆ −iθ ) ≤ p(xe−iθ ) =
|e−iθ |p(x) = p(x) [|e−iθ | = 1]. Hence | fˆ(x)| ≤ p(x) for all x ∈ X. Hence the result.
114
0.36. LINEAR FUNCTIONALS AND HAHN-BANACH THEOREM
fˆ X
= k f kZ , (0.36.3)
ˆ
| f (x)|
sup ≤ k f kZ .
x∈X,kxk=1 kxk
So f (x)ˆ ≤ k f kZ . But Z ⊂ X so ˆ
f (x) ≥ k f kZ . Hence ˆ
f (x) = k f kZ . Hence the
X X X
result.
Remark 0.36.1. In case of Hilbert space.
If Z is closed subspace of a Hilbert space X = H, then every bounded linear functional f has
a Riesz Representation, that is, f (x) = hx, Zi, Z is uniquely determined by f and kZk = k f k.
Since the inner product is defined on all of H so this gives at once a linear extension fˆ of f
from Z to H and fˆ has the same norm as f because fˆ = kZk = k f k.
115
0.37. APPLICATION OF HAHN BANACH THEOREM
Proof. We can find a non zero bounded functional f0 ∈ X 0 such that f0 (x) = kxk and k f0 k = 1.
Now,
| f (x)| | f0 (x)|
sup ≥ = kxk .
f 6=Θ k f k k f0 k
Again for all f 6= Θ ∈ X 0 , we have | f (x)| ≤ k f k kxk. Therefore,
| f (x)|
sup ≤ kxk .
f 6=Θ∈X 0 k f k
| f (x)|
Hence kxk = sup kfk
f 6=Θ∈X 0
116
0.37. APPLICATION OF HAHN BANACH THEOREM
Proof. Here f (x) = f (y), so f (x−y) = 0, for every f 6= Θ ∈ X 0 . Hence x−y = 0 and x = y.
Theorem 0.37.7. Let M be a closed subspace of a normed linear space X such that M 6= X.
Assume that u ∈ X\M and d = dist(u, M) = infm∈M ku − mk. Then there is a bounded linear
functional f ∈ X 0 such that
(i) f (x) = 0 for all x ∈ M.
(ii) f (u) = 1.
(iii) k f k = d1 .
Theorem 0.37.8. Let M be a subspace of a normed linear space X such that M 6= X. Assume
that u ∈ X\M such d = dist(u, M) = infm∈M ku − mk > 0. Then there is a bounded linear
functional f ∈ X 0 such that
(i) f (x) = 0 for all x ∈ M.
(ii) f (u) = d.
(iii) k f k = 1.
117
0.37. APPLICATION OF HAHN BANACH THEOREM
118
0.38. UNIFORM BOUNDEDNESS THEOREM
119
0.38. UNIFORM BOUNDEDNESS THEOREM
c
as n → ∞. Hence p ∈ Bm for every m. Since Bm ⊂ M m , we now see that p ∈ / Mm for every m,
so that p ∈
/ dMm = X. This contradicts that p ∈ X. Hence Baire’s theorem is proved.
Remark 0.38.1. The converse of Baire category theorem is not true. Infact there are incomplete
metric space which is nonmeager and shown by Bourbaki incomplete normed space which is
of second category.
From Baire’s theorem we shall now readily obtain the desired uniform boundedness
theorem. This theorem states that if X is a Banach space and a sequence of operators
Tn ∈ B(X,Y ) is bounded at every point x ∈ X, then the sequence is uniformly bounded. In
other words, pointwise boundedness implies boundedness in some stronger sense, namely,
uniform boundedness.
Theorem 0.38.2. Uniform Boundedness Theorem. Let (Tn ) be a sequence of bounded linear
operators Tn : X → Y , X is Banach space and Y is normed space such that (kTn xk) is bounded for
every x ∈ X, that is kTn xk ≤ cx , where cx is a real number depends on x. Then the sequence of
the norms kTn k is bounded, that is, there is a c(independent of x) such that kTn k ≤ c, n = 1, 2, ...
[pointwise boundedness of operator Tn : X → Y is uniformly bounded.]
Proof. For every k ∈ N, let Ak ⊂ X be the set of all x such that kTn xk ≤ k for all n.
Ak is closed. Indeed, for any x ∈ Ak there is a sequence (x j ) in Ak converging to x. This
means that for every fixed n we have Tn x j ≤ k and obtain kTn xk ≤ k because Tn is continuous
and so is the norm. Hence x ∈ Ak , and Ak is closed.
As (kTn xk) is bounded, so each x ∈ X belongs to some Ak . Hence X = ∪∞ k=1 Ak .
Since X is complete, Baire’s theorem implies that some Ak contains an open ball, say B0 =
B(x0 , r) ⊂ Ak0 .
r
Let x ∈ X be arbitrary not zero. We set z = x0 + γx, where γ = 2kxk . Then
kz − x0 k ≤ r,
so that z ∈ B0 . Hence from the definition of Ak and B0 = B(x0 , r) ⊂ Ak0 , we obtain kTn zk ≤ k0
for all n. Also kTn x0 k ≤ k0 since x0 ∈ B0 . Therefore we obtain x = z−x0
γ . For all n, we get
kTn (z − x0 )k 1 4
kTn xk = ≤ (kTn zk + kTn x0 k) ≤ kxk k0 .
γ γ r
Hence for all n,
4
kTn k = sup kTn xk ≤ k0
kxk=1 r
or kTn k ≤ c for c = 4r k0 .
Proposition 0.38.3. Space of Polynomials. The normed space X of all polynomials wlth norm
defined by
kxk = max|αi |,
i
where α0 , α1 , ... are coefficients of x, is not complete.
120
0.38. UNIFORM BOUNDEDNESS THEOREM
121
0.39. OPEN MAPPING THEOREM
Theorem 0.39.1. Open Mapping Theorem, Bounded Inverse Theorem. A bounded linear op-
erator T from a Banach space X onto a Banach space Y is an open mapping. Hence if T is
bijective, T −1 is continuous and thus bounded.
Lemma 0.39.2. (Open unit ball). A bounded linear operator T from a Banach space X onto a
Banach space Y has the property that the image T (B0 ) of the open unit ball B0 = B(0; 1) ⊂ X
contains an open ball about 0 ∈ Y .
122
0.39. OPEN MAPPING THEOREM
123
0.40. CLOSED GRAPH THEOREM
124
0.40. CLOSED GRAPH THEOREM
By definition, G(T ) is closed if and only if z = (x, y) ∈ G(T ) implies z ∈ G(T ). Also we
see that z ∈ G(T ) if and only if there are zn = (xn , T xn ) ∈ G(T ) such that zn → z, hence
xn → x, T xn → y,
and z = (x, y) ∈ G(T ) iff x ∈ D(T ) and y = T x. This proves the following useful criterion
which expresses a property that is often taken as a definition of closedness of a linear
operator.
Theorem 0.40.2. (Closed linear operator). Let T : D(T ) → Y be a linear operator, where
D(T ) ⊂ X and X and Y are normed spaces. Then T is closed if and only if it has the fol-
lowing property. If xn → x, where xn ∈ D(T ), and T xn → y, then x ∈ D(T ) and T x = y.
Note well that this property is different from the following property of a bounded
linear operator. If a linear operator T is bounded and thus continuous, and if (xn ) is a
sequence in D(T ) which conveIges in D(T ), then (T xn ) also converges.
This need not hold for a closed linear operator. However, if T is closed and two se-
quences (xn ) and (yn ) in the domain of T converge with the same limit and if the corre-
sponding sequences (T xn ) and (Tyn ) both converge, then the latter have the same limit.
Proof. Here T is not bounded as xn (t) = t n ,kxn k = 1 T xn (t) = xn0 = nt n−1 , so kT xn k = n and
kT xn k / kxn k = n, so there is no fixed c such that kT xn k / kxn k ≤ c.
Now, we’ll prove that T is closed by Closed linear operator theorem.
Let (xn ) in D(T ) be such that both (xn ) and (T xn ) converge, say, xn → x and T xn = xn0 → y.
Since convergence in the norm of C[O, 1] is uniform convergence on [0, 1], from xn0 → y,
we have Z t Z t Z t
0
y(τ)dτ = lim xn (τ)dτ = lim xn0 (τ)dτ = x(t) − x(0),
0 0 n→∞ n→∞ 0
that is, Z t
x(t) = x(0) + y(τ)dτ.
0
This shows that x ∈ D(T ) and x0 = y, because D(T ) ⊂ C[0, 1], set of all functions x which
has continuous derivatives. Hence T is closed.
It is worth noting that in above proposition, D(T ) is not closed in X since T would then
be bounded by the closed graph theorem.
Closedness does not imply boundedness of a linear operator. Conversely, boundedness
does not imply closedness.
For example: Let T : D(T ) → D(T ) ⊂ X be the identity operator on D(T ), where D(T )
is a proper dense subspace of a normed space X. Then it is trivial that T is linear and
bounded. However, T is not closed. This follows from closed linear operator theorem if
we take x ∈ X − D(T ) and a sequence (xn ) in D(T ) which converges to x.
Lemma 0.40.4. Closed Operator. Let T : D(T ) → Y be a bounded linear operator with domain
D(T ) ⊂ X, where X and Y are normed spaces. Then
(a) If D(T ) is a closed subset of X, then T is closed.
(b) If T is closed and Y is complete, then D(T ) is a closed subset of X.
125
0.40. CLOSED GRAPH THEOREM
Proof. (a) If (xn ) is in D(T ) and converges, say, xn → x, and is such that (T xn ) also converges,
then x ∈ D(T ) = D(T ) since D(T ) is closed, and T xn → T x since T is continuous. Hence T is
closed.
(b) For x ∈ D(T ), there is a sequence (xn ) in D(T ) such that xn → x. Since T is bounded,
kT xn − T xm k = kT (xn − xm )k ≤ kT k kxn − xm k .
126
0.41. PROJECTION OPERATORS
127
0.41. PROJECTION OPERATORS
Now (I − P)(P(x)) = P(x) − P2 (x) = 0, shows that null space of (I − P), that is, N(I − P)
satisfies Y ⊂ N(I − P). Again (I − P)(x) = 0 =⇒ x = P(x) ∈ Y . So N(I − P) ⊂ Y .
Consequently, N(I − P) = Y . So Y is always a closed subspace of H, as N(I − P) is a closed
subspace of H.
Finally P restricted to Y is the identity operator because y = P(x) gives P(y) = P(P(x)) =
P2 (x) = P(x) = y. This implies that P is a Projection operator on H.
Theorem 0.41.2. The projection operator P : H → H, H being a Hilbert space, is bounded and
self adjoint such that ||P|| = 1.
since hz1 , y2 i = 0. So hPx1 , x2 i = hx1 , Px2 i. Hence it follows that P is self adjoint.
128
0.42. COMPACT LINEAR OPERATORS
Theorem 0.42.3. Let (T n) be a sequence of compact linear operators from a normed space X
into a Banach spaces Y . If Tn → T (that is, ||Tn − T || → 0), then the limit operator T is compact.
Proof. Since T1 is a compact operator, we know that the sequence (T1 (xn )) has a convergent
(hence Cauchy) subsequence (T1 (x1,m )), where (x1,m ) is a subsequence of the original sequence
(xn ). The subsequence (x1,m ) is bounded, so we can repeat the argument with T2 to produce a
subsequence (x2,m ) of (x1,m ) with the property that (T2 (x2,m )) converges. We continue in the
same way, and then define a sequence (ym ) = (xm,m ). Notice that (ym ) is a subsequence of (xn ),
so it is bounded, say by ||yn || ≤ c, and it has the property that for every fixed n, the sequence
(Tn (ym )) is convergent, and hence Cauchy.
We claim that (T (ym )) is a Cauchy sequence in Y . Let ε > 0. Since ||Tn − T || → 0, there
ε
is some p ∈ N such that ||Tp − T || < 3c . Also, since (Tp (ym )) is Cauchy, there is some N > 0
such that ||Tp (y j ) − Tp (yk )|| < ε3 , whenever j, k > N. Therefore, for j, k > N, we have
Ty j − Tyk ≤ Ty j − Tp y j + Tp y j − Tp yk + Tp yk − Tyk
≤ Ty j − Tp y j + Tp y j − Tp yk + Tp yk − Tyk
ε
< T − Tp y j + + T − Tp kyk k
3
ε ε ε
< c + + c = ε.
3c 3 3c
Hence (T (ym )) is a Cauchy sequence in Y . Since Y is a Banach space, it is by definition
complete, so (T (ym )) converges. We have thus produced, for an arbitrary bounded sequence
(xn ) ⊂ X, a convergent subsequence of its image under T . Therefore, T is compact.
Example 0.42.4. Let T : `2 → `2 be an operator defined by T x = (αn xn ) for a sequence αn → 0.
We claim that T is compact. To show this, we approximate T by compact operators TN such
that TN x = (α1 x1 , . . . , αN xN , 0, . . .). As in the previous example we observe that TN is a compact
129
0.42. COMPACT LINEAR OPERATORS
This shows that ||TN − T || ≤ supn>N |αn |. Since αn → 0, it follows that ||TN − T || → 0. Hence
T is compact.
130
0.43. STRONG CONVERGENCE AND WEAK CONVERGENCE
131
0.43. STRONG CONVERGENCE AND WEAK CONVERGENCE
k n
(n)
||xn − x|| = ∑ α j e j − ∑ α je j
j=1 k=1
k
(n)
= ∑ (α j − α j )e j
j=1
k
(n)
≤ ∑ (α j − α j )e j
j=1
k
(n)
= ∑ |α j − α j| e j
j=1
ε
<k M = ε∀n ≥ N.
Mk
s
Thus xn →
− x.
Let X and Y be two normed linear space over the same field of scalars K.
Definition 0.43.3. Let {Tn } ∈ BLd (X,Y ) and let T ∈ BLd (X,Y ) then Tn → T in norm if for
every ε > 0 there exists a positive integer N such that ||Tn − T || < ε∀n ≥ N.
Definition 0.43.4. Let {Tn } ∈ BLd (X,Y ) and let T ∈ BLd (X,Y ) then Tn → T weakly written as
w w
Tn − → T x i.e. for every y0 ∈ Y 0 , y0 (T n(x) → y0 (T x).
→ T if for all x ∈ X, Tn x −
Definition 0.43.5. The sequence {Tn } ∈ BLd (X,Y ) is said to be strongly convergent if for every
x ∈ X, {Tn (x)} converges in Y .
Definition 0.43.6. Let {Tn } ∈ Bd L(X,Y ) and T ∈ Bd L(X,Y ). Then Tn is called strongly con-
s
vergent to T if for x ∈ X the sequence {Tn x} converges to T x and we write Tn →
− T.
Note that norm converges of a sequence of operators {Tn } ∈ BLd (X,Y ) to T ∈ BLd (X,Y )
implies its strong converges, because ||Tn (x) − T (x)|| = ||(Tn − T )x|| ≤ ||Tn − T ||||x|| → 0.
132
0.43. STRONG CONVERGENCE AND WEAK CONVERGENCE
Definition 0.43.7. Let {Tn } ∈ BLd (X,Y ) and let T ∈ BLd (X,Y ). Then {Tn x} converges to T (x)
uniformly for all x ∈ X such that ||x|| ≤ 1, if for every ε > 0, there exists a positive integer N
u
such that ||Tn x − T x|| < ε whenever n ≥ N written as Tn x →
− T x.
Alternatively we say uniform convergence of a sequence {Tn } ∈ BLd (X, y) in ||x|| ≤ 1
u
written as Tn →− T.
Theorem 0.43.5. If {Tn } ∈ BLd (X,Y ) be such that {Tn } ∈ BLd (X,Y ) uniformly for all x ∈ X
satisfying ||x|| ≤ 1 then Tn (x) → T (x) uniformly for all x ∈ X satisfying ||x|| ≤ 1.
Proof. Now, ||Tn x − T x|| = ||(Tn − T )(x)|| ≤ ||Tn − T ||||x||. Let x ∈ X such that ||x|| ≤ 1. So
u
||Tn x − T x|| → 0, implying Tn x →
− T x uniformly for all x ∈ X satisfying ||x|| ≤ 1.
u
Theorem 0.43.6. If {Tn } ∈ BLd (X,Y ) and T ∈ BLd (X,Y ) be such that Tn x →
− T x for all x ∈ X
such that ||x|| ≤ 1 then Tn → T in norm.
Proof. Let x ∈ X such that ||x|| ≤ 1. For any ε > 0 there exists N ∈ N such that ||Tn x − T x|| < ε
x
whenever n ≥ N. Let y = ||x|| , x 6= 0.
u
So, ||y|| = 1. So, Tn y →
− Ty, Therefore, we have
||Tn y − Ty|| < ε∀n ≥ N
||Tn x/||x|| − T x/||x|| || < ε∀n ≥ N
||Tn x − T x||/||x|| < ε∀n ≥ N
||(Tn − T )x||/||x|| < ε∀n ≥ N
||(Tn − T )x|| < ε||x||, ∀n ≥ N
So, sup ||x|| ≤ 1||(T n − T )x|| ≤ ε, ∀n ≥ N. So, ||(Tn − T )|| ≤ ε, ∀n ≥ N, implying that Tn → T
in norm.
s w
Theorem 0.43.7. Let {Tn } ∈ BLd (X,Y ) and let T ∈ BLd (X,Y ). If Tn →
− T then Tn −
→ T.
s
Proof. Let Tn → − T . Take ε > 0 be an arbitrary. So there exists a positive integer N such that
||Tn − T || < ε for all n ≥ N. Let y0 ∈ Y 0 . Take x ∈ X. Now,
|y0 (T nx) − y0 (T x)| = |y0 (Tn x − T x)|
≤ ||y0 ||||Tn x − T x||
≤ ||y0 ||||Tn − T ||||x|| → 0.
w w
Hence y0 (Tn x) → y0 (T x), Tn x −
→ T x, Tn −
→ T. Thus we have uniform convergence in ||x|| ≤
1.
So strong convergence implies weak convergence but weak convergence doesnot imply
its norm convergence, for example:
Example 0.43.8. Let X = `2 . Define Tn : X → X by Tn (x) = (0, 0, . . . , ξn+1 , ξn+2 , . . .), where
x = (ξ1 , ξ2 , . . . , ξn , . . .) ∈ X. Clearly for each positive integer n, Tn is linear. Now ||Tn (x)|| =
(∑∞ 2 1/2 ≤ ( ∞ |ξ |2 )1/2 = ||x||, implying that T ∈ BL (X,Y ). Since ∞ |ξ |2 < ∞
i=n+1 |ξi | ) ∑i=n i n d ∑i=1 i
w
so ξn → 0. So Tn (x) = (0, 0, . . . , 0, ξn+1 , ξn+2 , . . .) → (0, 0, . . . , 0, . . . , 0, . . .) = 0. Hence Tn −
→ Θ,
the zero operator on X. Take x = (0, 0, . . . , 0, ξn+1 , ξn+2 , . . .) ∈ X such that ||x|| ≤ 1. Then
Tn (x) = x. So ||Tn || = 1 for all n. But ||Tn − Θ|| = ||Tn || = 1, implying that Tn 9 Θ in norm.
So by {Tn } does not converge uniform to the zero operator.
133
0.44. REFERENCES
0.44 References
1. Kreyszig, E., Introductory Functional Analysis with Applications, Wiley (1989).
3. Simmons, G.F., Introduction to Topology and Modern Analysis, Tata McGraw Hill (1963).
134