Functional Notes-2
Functional Notes-2
Dr. Deepesh K. P.
linear spaces
The whole mathematics is developed from sets. One can proceed to study mathematics from sets by
forming collection of subsets of a given set and develop theories (Topology and σ-algebras of measure
theory). Another way is to introduce an operations on a given set. If we study about a set by introducing
one suitable operation between elements of the set, it can lead to Group theory. If more than one
operations are considered between elements of the given set, we can think of developing Ring theory. If
the second operation is performed between an element of the set with a suitable collection of scalars, then
it leads to Linear algebra which is built upon the vector space structure.
In functional analysis, the basic building materials used are vector spaces. First let us recollect some
of the basic ideas from Linear algebra. Then onwards we will be developing all concepts and try to be
self contained as long as possible (except for some basic Real analysis and Topology concepts).
Definition 1.0.1. A non empty set V together with two operations namely the vector addition ‘+′ and
the scalar multiplication ‘.′ is called a vector space if it satisfy the following properties (here we assume
1. u + v ∈ V
2. u + v = v + u
3. u + (v + w) = (u + v) + w
1
2
6. α.u ∈ V
10. 1.u = u
The elements of a vector space V are called vectors and the real numbers α, β, . . . are called scalars.
For the purpose of discussing functional analysis, we consider the scalars to be either from the set of
all real numbers R or from the set of all compled numbers C. The scalar field is denoted by K in this
discussion (this means K represents either R or C). A set should satisfy all the 10 properties of the
operations if it is to be called a vector space. If any one of these 10 properties is not satisfied, the set
An immediate way to get more vector spaces is to look inside the vector space and search for suitable
This means that a vector space sitting inside a bigger vector space is called a subspace. Of course,
the operations in the subspace should be the same as that of the bigger space.
It is easy to observe that a subset qualifies to be a subspace of the vector space if it satisfies the
properties (1) and (6) of the definition of vector space. That is, a subset W ⊆ V if and only if w1 +w2 ∈ W
Example 1.0.3. Let S be any set. Let us call F(S, K) to be the collection of all functions from S to K.
That is,
F(S, K) = {f : S → K is a function}.
for f, g ∈ F(S, K) and α ∈ K. Then since the images of the functions are from R or C, and these spaces
have the associative, commutative, distributive and all other properties, we can show that F(S, K) under
3
the two operations forms a vector space. Note that when S = {1, 2, . . . , n}, we get the collection of all
n−tuples, when S = N, we get the collection of all real/complex sequences and when S is an interval
like [a, b] in R, we get the standard function spaces like the polynomials P[a, b] and continuous function
In a vector space, from a given set of elements, we can create more and more dependent elements by
u = α1 u1 + α2 u2 + . . . + αn un
is called a linear combination of the vectors u1 , u2 , . . . , un . Note that u ∈ V . The collection of all possible
finite linear combinations of elements of a set S is called the linear span of the set, denoted as span(S).
Definition 1.0.5. If E ⊂ V is an infinite set then the span of E is the collection of all finite linear
We say that a subset of V is linearly dependent if at least one of the elements can be constructed
Theorem 1.0.7. A collection of vectors {u1 , u2 , . . . , un } in V is linearly dependent if and only if there
α1 u1 + α2 u2 + . . . + αn un = 0.
Sets which are not linearly dependent are called linearly independent. They can be independently
if and only if
α1 u1 + α2 u2 + . . . + αn un = 0 ⇒ α1 = 0, α2 = 0, . . . , αn = 0
An infinite set E ⊂ V is said to be linearly independent if each finite subset of E is linearly independent.
There are some subsets of V which can span the entire vector space(create the entire vector space by
Definition 1.0.9. A subset S of a vector space V is said to be a spanning set if every element of V can
Note that this S can be very large and any super set of a spanning set is again a spanning set.
Note that the spanning sets may contain linearly dependent elements, which are actually useless (as some
of them can be created using others). So avoiding the linearly dependent elements from a spanning set
Definition 1.0.10. A linearly independent subset which spans the vector space is called a basis (or Hamel
basis) of the vector space. That is, a collection of vectors B is a basis of V if and only if it is linearly
Every vector of V can be written as a linear combination of elements of a spanning set. When the
Theorem 1.0.11. Let B be a basis for a vector space V . Then for each v ∈ V , there exists a unique
v = c1 v1 + c2 v2 + . . . + cn vn .
In many cases, we identify the vector v with the scalar set (c1 , c2 , . . . , cn ) in linear algebra. Note that
infinite dimensional spaces also can have good bases. For example {1, x, x2 . . .} form a countable basis
for P, the collection of all polynomials on R and the collection of vectors {e1 , e2 , . . .} with
en = (0, 0, . . . , 0, 1, 0, . . .),
where 1 appears in the nth coordinate, for each n ∈ N, forms a countable basis for c00 , the space of all
sequences which are vanishing after a finite number of terms. We will see details of this space later.
Even though there can be infinitely many different bases for a vector space, all of them have a unique
Theorem 1.0.12. The cardinality of each basis of a vector space will be the same. This number is called
The basic question one may feel to ask is about the need of a different structure on a vector space as it
already has lot of operations on it. Note that Linear algebra mostly deals with finite dimensional vector
spaces (of course, infinite dimensional spaces can be discussed with the two operations also) and linear
transformations between them which can be represented by m × n matrices (which is not possible when
you have an infinite dimensional vector space, in general). Further, even though we can talk about linear
combinations in a vector space, infinite linear combinations not possible to be treated in a vector space
structure if we do not have concept of convergence of sequences. To talk about convergence of sequences
you need a metric structure or, at least, a topological structure on a vector space.
Even though we talk about linear transformations, which are particular functions between vector
spaces, we can not talk about the continuity of functions in the settings of linear algebra. Further, to talk
about closed-ness of elements or to discuss completeness property and all, we need a metric to be defined
on a vector space. Some of the other interesting concepts that one would like to discuss for vectors in a
vector space are the concepts of length of vectors and geometry of vectors. For dealing with such things,
We will see that inner products induces a norm; norm induces a metric and a metric induces a topology
on vector spaces. Hence the concept of topological vector spaces is the most general among these. But
in our Functional analysis we start our study with the in between concept, that is the concept of normed
linear spaces as the theory is very strong and considerably general in this settings, which allows us to
Definition 1.0.13. A norm on a vector space X is a function ∥ · ∥ : X → R such that for any x, y ∈ X
2. ∥λx∥ = |λ|∥x∥
6
3. ∥x + y∥ ≤ ∥x∥ + ∥y∥
If a linear space X is equipped with a norm ∥ · ∥, we call (X, ∥ · ∥) a normed linear space.
We have seen that F(S, K) is a vector space and when S = {1, 2, . . . , n}, we get Kn = {(x1 , x2 , . . . , xn ) :
is a normed linear space. The property |a + b| ≤ |a| + |b| of complex numbers helps to verify the norm
is also a norm on Kn .
When S = N, F(S, K) becomes the collection of all scalar sequences. It is not difficult to see that the
previous norms, when extended, forms norms on certain subspaces of F(S, K). That is, the space
∞
X
ℓ1 (N) = {(an ) : an ∈ K and |an | < ∞}
n=1
forms a norm on ℓ1 (N) (we write this space as ℓ1 ). Again, the collection
∥x + y∥ ≤ C(∥x∥ + ∥y∥)
for some C > 0, for all x, y ∈ X, then the quantity ∥ · ∥ is called a quasinorm on X.
Definition 1.2.2. If the condition ”∥x∥ = 0 iff x = 0” is omitted from Definition 1.0.13, then the
We discuss some of the properties of norm here. These can be easily proved. First we give a refined
Proof. This follows from the triangle inequality applied to ∥x∥ = ∥(x − y) + y∥.
Proof. Use the fact |∥x∥ − ∥y∥| ≤ ∥x − y∥. So if we take f : X → R by f (x) = ∥x∥ for x ∈ X, for any
ϵ > 0 we can choose δ = ϵ such that |f (x) − f (y)| < ϵ whenever ∥x − y∥ < δ.
It is easy to see that every norm induces a metric on a normed linear space.
d(x, y) = ∥x − y∥,
then d is a metric on X.
Thus every normed linear space is a metric space, and therefore a topological space also. What are
Definition 1.2.6. Let X be a normed linear space and x0 ∈ X. Then the open ball with center at x0
We call B(0, 1) as the open unit ball of X. The collection of all open balls form a basis for the metric
topology, which is the topology on X. As in the case of a metric space, the convergence of a sequence is
Proof. Take x, y ∈ B(x0 , r) and show that for any 0 < t < 1, tx + (1 − t)y ∈ B(x0 , r).
Theorem 1.2.9. Suppose (X, ∥ · ∥) is a normed linear space and Y be a subspace of X. Then (Y, ∥ · ∥)
Definition 1.2.10. Consider the following subsets of the sequence space F(N, K).
2. c0 = {(an ) : an = 0 as n → ∞}
Theorem 1.2.11. The spaces c, c0 , c00 with ∥ · ∥∞ are normed linear spaces.
Proof. In view of Theorem 1.2.9, it is enough to show that these are subspaces of ℓ∞ .
There are some standard spaces used in functional analysis to develop the useful structures and theory
and also to provide examples for illustrating/disproving results. We will see these spaces in this section.
We have already seen that (Kn , ∥ · ∥1 ) and (Kn with ∥ · ∥∞ ) are normed linear spaces. We can define
certain other norms on this space which are useful and gives more examples of normed linear spaces.
9
n
! p1
1 X
∥(x1 , x2 , . . . , xn )∥p = (|x1 |p + |x2 |p + . . . + |xn |p ) = p |xi |p .
i=1
We will show that ∥ · ∥p is a norm on Kn . For this we need some famous inequalities.
Young’s inequality
a+b
This is a generalization of the AM-GM inequality. From (a − b)2 ≥ 0, we get 2 ≥ ab. We can re-write
this as
a2 b2
+ ≥ ab
2 2
by taking a2 and b2 in place of a and b. A generalized version of this is called the Young’s ineqaulity.
1 1
Theorem 1.4.1. (Young’s inequality) Let a, b ≥ 0, 1 < p < ∞ and 1 < q < ∞ be such that p + q = 1.
Then
ap bq
+ ≥ ab.
p q
a ap 1
Proof. Instead, we will show that − − ≤ 0 (just by dividing with bq throughout), assuming
bq−1 pbq q
1 1 1 ap
b ̸= 0, since if b = 0, it is trivial. Make a function f : [0, ∞) → R by f (t) = t p − t − by taking t = q .
p q b
Now show that f is increasing on the left side of 1 and decreasing to the right side of 1, showing that the
maximum of f is at t = 1, which is 0.
This inequality leads to the famous Holder’s inequality, which is very useful in proving the triangle
Let (an )n1 , (bn )n1 be n-tuples in Kn and p, q be Holder’s pairs. Then
n n
! p1 n
! q1
X X X
p q
|ai bi | ≤ |ai | |bi |
i=1 i=1 i=1
Let (an )∞ ∞
1 , (bn )1 be sequences in K and p, q be Holder’s pairs. Then
∞ ∞
! p1 ∞
! q1
X X X
|ai bi | ≤ |ai |p |bi |q
i=1 i=1 i=1
10
proof: If unbounded, nothing to show. For bounded, show that partial sums are bounded by.. Note
Once we have the Holder’s inequality, we can get the Minkowski’s inequality which is nothing but the
Let (an )n1 , (bn )n1 be n-tuples in Kn and 1 < p < ∞. Then
n
! p1 n
! p1 n
! p1
X X X
p p p
|ai + bi | ≤ |ai | + |bi |
i=1 i=1 i=1
∞
! p1 ∞
! p1 ∞
! p1
X X X
|ai + bi |p ≤ |ai |p + |bi |p
i=1 i=1 i=1
These results helps us to prove the norm properties for some of the standard spaces.
proof.
Proof:
Note that c00 being a subspace of ℓp is a normed linear space with ∥ · ∥p ; whereas c0 can’t be defined with
Theorem 1.4.9. Suppose 1 ≤ p < q ≤ ∞. Then ℓp ⊂ ℓq . Show that the inclusion is proper.
Proof. For q = ∞, it follows from the fact that convergent sequences are bounded. For 1 ≤ p < q < ∞,
We know that F(S, K) is a vector space for any set S. In general, we cant define a norm on this vector
space. But there are useful cases and subspaces of this vector space other than Kn and sequence spaces,
11
Then this is a norm (easy!) and it makes ℓ∞ (S) a normed linear space. For this we do not need any
additional structure on S. This is totally different from the Lp (E) space which is going to come soon.
Consider
Since C[a, b] is a subspace of ℓ∞ ([a, b]), it follows that C[a, b] is a normed linear space with ∥ · ∥∞ . The
same is true for P[a, b] and C ′ [a, b] - space of all smooth functions- also.
Note that one can replace [a, b] with any compact metric space Ω also. This is denoted as C(Ω), which
Z b
∥f ∥1 = |f (t)|dt, for f ∈ C[a, b],
a
where the integral is the usual Riemann integral. It is easy to verify the norm properties.
Like in the case of sequences, we can generalize the 1−integral to p−integrals for 1 < p < ∞ and derive
! p1
Z b
p
∥f ∥p = |f (t)| dt , for f ∈ C[a, b]
a
is a norm on C[a, b] for 1 < p < ∞. To show this we need the Holder’s and Minkowski’s inequalities for
1 1
Theorem 1.7.1. Let 1 < p < ∞ and q be such that p + q = 1. Then for any f, g ∈ C[a, b],
! p1 ! q1
Z b Z b Z b
p q
|f (t) + g(t)|dt ≤ |f (t)| dt |g(t)| dt
a a a
! p1 ! p1 ! p1
Z b Z b Z b
|f (t) + g(t)|p dt ≤ |f (t)|p dt + |g(t)|p dt
a a a
This essentially gives the triangle inequality which can be used to prove that ∥ · ∥p is a norm on C[a, b].
Theorem 1.7.4. The spaces P[a, b], C ′ [a, b] with ∥ · ∥p , 1 < p < ∞ is a normed linear space.
We have seen that C[a, b] is a normed linear space with ∥ · ∥p , 1 ≤ p ≤ ∞. But there are a lot of
useful functions which are not continuous. Further, since the above p-norms are using integrability
and integrable functions are more general in the sense that they do not need stringent conditions like
continuity or differentiablity, we define certain subclasses of functions space which can be defined using
4. Show that every semi-norm on a vectorspace induces a norm in a suitable quotient space.
8. Show that if p > 0 and ∥.∥ is a norm on a vectorspace X, ∥x∥∗ = p∥x∥ is also a norm on X.
13
9. Show that the distance induced by a norm on a nonzero normed space can never be bounded.
14. Show that (Kn , ∥.∥p ) is a normed linear space for 1 < p < ∞.
15. Show that (ℓp , ∥.∥p ) is a normed linear space for 1 < p < ∞.
17. If 1 ≤ p < q < ∞, show that ∥an ∥q ≤ ∥an ∥p , for any sequence (an ) in ℓp .
18. For 1 ≤ p < q ≤ ∞, show that ℓp ⊂ ℓq . Show that the containment is strict.
∞
p1
p p
P
19. If 0 < p < 1, then ℓ is a quasi-normed linear space with ∥an ∥p = |an | .
n=1
1
Hint: Use (a + b)p ≤ ap + bp for 0 < p < 1 and f (x) = x p is a convex function for such p.
21. Show that Lp (R) and Lq (R) are not comparable, 1 ≤ p < q ≤ ∞ (see that µ(R) = ∞)
25. Show that if (X1 , ∥ · ∥1 ) and (X2 , ∥ · ∥2 ) are normed spaces, then the cartesian product X1 × X2 is
also a normed linear space with the norm ∥(x, y)∥ = ∥x∥1 + ∥y∥2 for (x, y) ∈ X1 × X2 .
14
26. Let V be the vectorspace of all continuous complex valued functions on J = [a, b]. Let X1 =
Z b
1/2
∥x∥2 = ⟨x, x⟩ and ⟨x, y⟩ = x(t)y(t)dt.
a
27. Orthonormalize the first three terms of the sequence (x0 , x1 , x2 , . . .), where xi (t) = ti , on the interval
[-1,1], where
Z 1
⟨x, y⟩ = x(t)y(t)dt.
−1
Chapter 2
Suppose V is a vector space on K, where K stands for the set of all real or complex numbers. In this
c. The space Pn [a, b], the space of all complex polynomials of degree atmost n with domain [a, b].
d. The space P[a, b], the space of all complex polynomials with domain [a, b].
e. The space C[a, b], the space of all complex valued functions with domain [a, b].
f. The space L2 [a, b], the space of all 2-integrable complex functions with domain [a, b].
An inner product on a vector space helps to introduce certain geometry to the vector space (like the
orthogonality). It can be also thought as an extension of the concept of dot product of 3-dimensional
Definition 2.0.1. A vector space V is said to be an inner product space if there is defined a function
15
16
The above defined function ⟨, ⟩ on V is called an inner product on V (note that on all vector spaces,
we may not be able to define an inner product). The properties 1 and 2 are combinedly called positive
definiteness and the properties 3 and 4 together are called linearity properties. Property 5 is called
conjugate symmetry. Note that property 5 becomes ⟨u, v⟩ = ⟨v, u⟩ if the inner product is real.
We use the notation (X, ⟨·, ·⟩) for an inner product space (instead of V) and the elements of the space
Example 2.1.1. 1. Take X = R and define ⟨x, y⟩ = xy, the usual multiplication of real numbers.
n
X
n
3. Take X = K and define ⟨(x1 , x2 , . . . xn ), (y1 , y2 , . . . yn )⟩ = xj yj
j=1
∞
X
4. Take X = ℓ2 and define ⟨(x1 , x2 , . . .), (y1 , y2 , . . .)⟩ = xj yj
j=1
Z b
5. Take X = Pn [a, b] and define ⟨p(x), q(x)⟩ = p(x)q(x)dx
a
Z b
6. Take X = C[a, b] and define ⟨f, g⟩ = f (x)g(x)dx
a
Z b
2
7. Take X = L [a, b] and define ⟨f, g⟩ = f (x)g(x)dµ(x)
a
Here as an illustration, we verify the inner product properties for 6 only. Here
Z b
⟨f, g⟩ = f (x)g(x)dx.
a
Z b
a. ⟨f, f ⟩ = |f (x)|2 dx ≥ 0 since the integral of a non negative function over any subset of R is non
a
negative.
Z b
b. ⟨f, f ⟩ = |f (x)|2 dx = 0 if and only if f = 0, the zero function.
a
Z b Z b Z b
c. ⟨f + h, g⟩ = (f (x) + h(x))g(x)dx = f (x)g(x)dx + h(x)g(x)dx = ⟨f, g⟩ + ⟨h, g⟩
a a a
Z b Z b
d. ⟨f, g⟩ = f (x)g(x)dx = g(x)f (x)dx = ⟨g, f ⟩
a a
17
We can see that the inner product induces a norm on the inner product space by
p
∥x∥ = ⟨x, x⟩, ∀x ∈ X.
But in order to show that this positive quantity gives a norm on the inner product space we need a result
Theorem 2.2.1. Suppose X is an inner product space and x, y ∈ X. Then these vectors satisfy the so
Proof. Start with ∥x − αy∥2 = ⟨x − αy, x − αy⟩, which is a positive quantity. Now expand it to get
⟨x,y⟩
∥x∥2 − 2Re(⟨x, αy⟩) + |α|2 ∥y∥2 ≥ 0. Since it is true for any α ∈ C, choosing α = ∥y∥2 , for y ̸= 0, we get
the conclusion.
Note 2.2.2. Note that Schwartz inequality becomes the Holder’s inequality (with p = 2) in the case
of n−tuples, sequences or L2 functions. Once we have the Schwartz inequality, we can show that ∥x∥
⟨x, y⟩
⟨x, y⟩ = ∥x∥∥y∥ cos θ ⇒ p p = cos θ
⟨x, x⟩ ⟨y, y⟩
and | cos θ| ≤ 1 for all θ, resulting in |⟨x, y⟩| ≤ ∥x∥ ∥y∥. The Schwartz inequality says that this is true
Except the triangle inequality, other properties easily gets followed from the properties of inner product.
18
As we have seen norm is a continuous function. We see that inner products are also continuous in the
individual variables. That is, for a fixed y ∈ X, if we define f (x) = ⟨x, y⟩, then f : X → K is continuous.
Theorem 2.2.5. Inner product is continuous in the first and second variable. This means if xn → x as
n → ∞, then for each y ∈ X fixed, ⟨xn , y⟩ → ⟨x, y⟩ (also ⟨y, xn ⟩ → ⟨y, x⟩).
Proof. Use the idea: |⟨xn , y⟩ − ⟨x, y⟩| = |⟨xn − x, y⟩| ≤ ∥xn − x∥∥y∥ by Schwartz inequality.
The above norm is called the norm induced by the inner product. This norm induces a metric or
d(x, y) = ∥x − y∥ ∀x, y ∈ X,
giving it a metric structure. Thus every inner product space is a normed space and every normed space
Note that there are metric spaces which are not normed spaces (eg. any bounded metric space cannot
be a normed space!) and there are normed spaces which are not inner product spaces. For example we
To show that the norm of a particular normed space is not induced by any inner product, we should
know the properties of the norms created from inner products. Such a property is the Parallelogram law.
Theorem 2.3.1. Suppose X is an inner product space and x, y ∈ X. Then these vectors satisfy the so
Proof. Start with ∥x + y∥2 + ∥x − y∥2 = ⟨x + y, x + y⟩ + ⟨x − y, x − y⟩ and cancel out the terms ⟨x, y⟩
This is called the parallelogram identity because in the case of 2-dimensional vectors, if two vectors
are represented as the two adjacent sides of a parallelogram, then the diagonals represents the sum and
difference of the vectors. In a parallelogram it is true that the sum of the squares of the diagonal lengths
In other words, among the norms ∥∥p , only the norm ∥∥2 is induced by an inner product.
Proof. We already know the inner product on ℓ2 induces the ∥∥2 . Now show that parallelogram identity
Note that the same is true for (Cn , ∥∥p ) using the same technique.
We know that from an inner product, a norm arises and the expression for the norm in terms of the
p
inner product is ∥x∥ = ⟨x, x⟩. A natural question is, if we know that a norm is induced from an inner
Theorem 2.3.3. (Polarization identity) Suppose (X, ⟨ ⟩) is an inner product space and ∥ · ∥ be the norm
induced. Then
1
∥x + y∥2 − ∥x − y∥2 + i(∥x + iy∥2 − ∥x − iy∥2 )
⟨x, y⟩ =
4
and the imaginary part is not required if the space is a real inner product space.
Note 2.3.4. An information: Suppose we have a normed linear space (X, ∥ · ∥). Is there any criterian to
check if this norm is induced by some inner product? The classical result by Von Nuemann says that a
norm is induced by an inner product if and only if the norm satisfies the Parallelogram identity. In that
case the Polarization identity helps us to find the inner product from the norm.
Definition 2.4.1. We say that two elements x, y ∈ X are orthogonal to each other if ⟨x, y⟩ = 0. We say
Example 2.4.2. The vectors (1, 0, 1) and (1, 0, −1) are orthogonal in C3 and sin x; and cos x are or-
thogonal in C[−1, 1]. The set {(1, 0, 0, . . .), (0, 1, 0, . . .), (0, 0, 1, . . .), . . .} is an orthogonal set in ℓ2 .
20
Definition 2.4.3. We say that a set E in X is an orthonormal set if it is an orthogonal set and all
√
Theorem 2.4.4. Show that any two different elements of an orthonormal set are at a distance of 2.
Example 2.4.5. The vectors (1, 0, 1) and (1, 0, −1) are orthogonal in C3 but not orthonormal ; Same
is the case with sin x and cos x in C[−1, 1]. The set {(1, 0, 0, . . .), (0, 1, 0, . . .), (0, 0, 1, . . .), . . .} is an
orthonormal set in ℓ2 .
Note 2.4.6. Note that if E = {xα }α∈A is an orthogonal set (where A is some index set) which does not
xα
contain 0, then is an orthonormal set in X.
∥xα ∥
Another geometrical aspect of inner product spaces is that vectors which are orthogonal to each other
Theorem 2.4.7. Suppose X is an inner product space and x, y are orthogonal in X. Then these vectors
Note 2.4.8. This result can be extended to n vectors. That is if x1 , x2 , . . . , xn are orthogonal in an inner
What is the connection between orthogonal sets and linearly independent sets?
But linearly independent sets need not be orthonormal (or even orthogonal)!. However, we can show
that from every sequence of linearly independent set, we can generate an orthonormal set.
21
orthonormalization process:
u1 = y1
⟨y2 , u1 ⟩
u2 = y2 − u1
∥u1 ∥2
⟨y3 , u1 ⟩ ⟨y3 , u2 ⟩
u3 = y3 − u1 − u2
∥u1 ∥2 ∥u2 ∥2
......
n−1
X ⟨yn , ui ⟩
un = yn − ui
i=1
∥ui ∥2
......
u1 u2 u3
one can create the set {u1 , u2 , u3 , . . .}, which will be an orthogonal set and the set , , ,...
∥u1 ∥ ∥u2 ∥ ∥u3 ∥
will be an orthonormal set in X. Further the span of {y1 , y2 , . . . , yn } and {u1 , u2 , u3 , . . . , un } are equal
for each n ∈ N.
Proof. It is a verification. For the last part, observe that y1 , y2 , . . . , yn ∈ span{u1 , u2 , . . . , un } and
We have seen that two elements x, y are orthogonal (x ⊥ y) if ⟨x, y⟩ = 0. In the following we define
Definition 2.4.11. For a subset A of an inner product space X, the orthogonal compliment (or annihi-
lator) of A is denoted by A⊥ and is defined as the set of all those elements which are orthogonal to every
element of A. That is
It can be seen that from each non empty subset A, we can create a closed subspace in the inner
product space.
Theorem 2.4.12. Let A be a subset of an inner product space X. Then A⊥ is a closed subspace of X.
Proof. Show the sums and scalar multiples are inside to show that it is subspace. To show closedness,
take a sequence (xn ) in A⊥ which converges to x in X. To show that x ∈ A⊥ , use the fact that the inner
2.5 Problems
6. Show that equality happens in Schawrtz inequality iff x and y are dependent.
|⟨x,αy⟩)|
Hint: ∥x − αy∥2 = ∥x∥2 − ∥y∥2
7. Show that in a real inner product space, ∥x + y∥2 = ∥x∥2 + ∥y∥2 if and only if x ⊥ y.
Hint: One side is true for every inner product space (Pythagorus Theorem). For the other side,
expand ∥x + y∥2 = ∥x∥2 + ∥y∥2 using inner product and cancel out ⟨x, x⟩ and ⟨y, y⟩. Now use
Hint: Assume ∥x∥2 = ∥y∥2 . Expand ⟨x + y, x − y⟩ and cancel out ∥x∥2 and ∥y∥2 . In the remaining
Hint: Squaring ⟨x + αy, x + αy⟩ = ⟨x − αy, x − αy⟩. Expand and cancel the terms ⟨x, x⟩ and
|α|2 ⟨y, y⟩, resulting in 2{α⟨y, x⟩ + α⟨x, y⟩} = 0 . That is Real(α⟨y, x⟩) = 0. Choose α = 1 and then
10. Show that {(1, 2), (1, 3)} is a linearly independent set in the inner product space R2 , whereas they
Hint: 0 is the one and only one element orthogonal to all elements of X.
23
Hint: Take any two elements u, v ∈ A⊥ and α ∈ K and show that ⟨u + αv, x⟩ = 0 for all x ∈ A.
16. If A ⊂ B in X, then B ⊥ ⊂ A⊥ .
Hint: Let x ∈ B ⊥ . Then ⟨x, y⟩ = 0 for all y ∈ B. But then ⟨x, z⟩ = 0 for all z ∈ A, since A ⊂ B.
17. If A ⊂ X, then A⊥ = A⊥ .
Soln: A⊥ is closed.
⊥
18. If A ⊂ X, then A⊥ = A .
⊥
Hint: Since A ⊂ A, by [5], A ⊂ A⊥ . For converse, take x ∈ A⊥ . Then ⟨x, y⟩ = 0 for all y ∈ A. To
⊥
show x ∈ A , let z ∈ A. Then z = lim zn , for zn ∈ A. Now use ⟨x, z⟩ = ⟨x, lim zn ⟩ = 0.
n→∞ n→∞
⊥
Hint: S ⊥ = S = X ⊥ = 0.
We first discuss about convergent sequences and Cauchy sequences in a metric space.
Theorem 3.0.1. If ∥ · ∥1 and ∥ · ∥2 are two equivalent norms on a vector space X. Then (X, ∥ · ∥1 ) is
So if KN with ∥∥∞ is a Banach space, then KN with ∥∥1 is also a Banach space.
We have seen that KN with ∥∥∞ is a Banach space and therefore KN with ∥∥1 is also a Banach space. We
will see that any finite dimensional normed space is a Banach space. We also give some characterizations
First let us define some norms which can be defined on every finite dimensional spaces.
25
26
Proof. Same as the proof of KN . Take a Cauchy sequence, show that the coordinate sequences are Cauchy
and use their limits to define the limit of the Cauchy sequence.
Definition 3.1.3. Suppose (X, d) is a metric space and A ⊂ X. For a point x0 ∈ X, we define the
distance from x0 to A as
Now we want to show that a finite dimensional normed linear space is complete w. r. to any norm
Theorem 3.1.4. Any two norms are equivalent on a finite dimensional normed linear space .
Proof. Let X be a normed linear space with basis {u1 , u2 , . . . , uN } and ∥∥ be any norm on X. We show
that ∥ · ∥ is equivalent to ∥∥∞ . Now using the transitivity of equivalent norms, we get the conclusion.
N
X N
X
It is easy to see that ∥x∥ = ∥ αi ui ∥ ≤ ∥x∥∞ ∥ui ∥.
i=1 i=1
For the converse part, observe that
N
X
∥x∥ = ∥ α i ui ∥
i=1
≥ |αi |dist(ui , Yi )
where Yi = span{u1 , . . . , ui−1 , ui+1 , . . . , uN }. Note that mini=1,2,...,N dist(ui , Yi ) > 0. Now taking supre-
∥x∥ ≥ m0 ∥x∥∞ ,
This, along with Theorem 3.1.2, immediately gives the following result.
Theorem 3.1.5. Every finite-dimensional normed linear space is complete w. r. to any norm.
Proof. Any two norms on finite dimensional normed linear space are equivalent. So if the space is complete
w. r. to one of the norms, it will be complete w.r.to any other norm also.
27
So finite dimensional spaces are very good examples of Banach spaces. But how do we identify if a
normed linear space is finite-dimensional, if we do not have information about the basis? Now we obtain
a characterization for finite-dimensional normed linear space in terms of the topological properties of
their closed unit balls. For this purpose, we first quote a famous result from the theory of metric spaces.
Theorem 3.1.6. Suppose (X, d) is any metric space and A ⊂ X. Then T.F.A.E:
i. A is compact
Corollary 3.1.7. Suppose (X, d) is a complete metric space and A ⊂ X. Then A is compact if and only
For the space K, we can see that all bounded sets are totally bounded also (using the idea that subsets
of totally bounded sets are totally bounded). So in K, subsets are bounded if and only if they are totally
Theorem 3.1.8. A subset of K is compact if and only if it is closed and totally bounded.
Not only for K, the above result is true for any finite dimensional normed linear spaces, w. r. to any
norm.
A subset of a finite dimensional normed linear space is compact if and only if it is closed and bounded.
Proof. Let A ⊂ X be compact. Then by Theorem 3.1.6, A is complete and totally bounded. Hence A is
Conversely, suppose A is closed and bounded. Let us take a basis {u1 , u2 , . . . , uN } be a basis for X.
We will show that every sequence in A has a convergent subsequence. Take (xn ) in A, which is a bounded
set. Then
∥xn ∥∞ ∼ ∥xn ∥ ≤ M.
X
If xn = αn(j) uj , this means |αnj | ≤ M for j ∈ {1, 2, . . . , N }. Now find subsequence αnj k → αj . Now
X
form x = α(j) uj and show that xnk → x as k → ∞. Now x ∈ A, being closed.
Thus the closed unit ball is compact in a finite dimensional normed linear space . But if the space
is infinite dimensional this can’t happen. We first give an example and then show that this is actually a
characterizing property of finite dimensional spaces with the help of Riesz Lemma.
28
Example 3.1.10. Consider ℓ∞ . Then the closed unit ball contains en = (0, 0, . . . , 0, 1, 0 . . .), where 1 is
at the nth place. Then ∥en − em ∥ = 1 for each n, m ∈ N. Hence closed unit ball is not a compact set.
Next we define a fundamental result on normed linear space , called the Riesz Lemma.
Let Y be a closed proper subspace of a normed linear space X and 0 < r < 1. Then there exists an
Proof. Use
We will see that many of the standard spaces are Banach spaces. If a space is not Banach, then it is
always possible to find a Banach space which contains this space as a dense subset. This constructive
proof we are not discussing here. Now we give some examples of standard infinite dimensional (all finite
Proof. Write
Proof. Write
Proof. Write
Proof. Write
Proof. Write
Proof. Write
Proof. Write
We show another useful characterization of Banach spaces. For this, we introduce some concepts.
The following property helps us to show that some normed linear space s are not Banach spaces.
Theorem 3.2.10. A normed linear space X is a Banach space iff every absolutely summable series are
summable.
Example: In c00 , we have absolutely summable series ( n12 ), which is not summable. Hence it is ???
Now we study the relationship between completeness and basis. We have seen that if a normed linear
space has a finite basis, then it must be complete w. r. to any norm Theorem 3.1.5. Now, what if
the basis is infinite? Can we conclude something? Generally, the answer is no since there+ are infinite
dimensional spaces like c00 which is incomplete in ∥∥∞ and ℓ1 , which is complete. But if the basis is
denumerable (infinite, but countable), then we can conclude that the space can never be a Banach space.
To do this, we need a result on metric spaces, namely, the Baire category theorem. We first define some
0
Definition 3.3.1. A subset A of a metric space X is said to be nowhere-dense subset if A = ϕ.
This means A contains no interior points. It should be observed that, in such a case, not only A but
What is the connection between dense and nowhere-dense sets? They are strongly complementary in
0
some sense. That is if A is dense in X, then A = X and so A also equals the whole space X, whereas
nowhere dense sets have the opposite. Also, we know that if A is dense in X, its complement does not
contain any open ball inside; whereas if A is nowhere dense, A and A do not contain open balls.
30
Examples of nowhere-dense sets are finite sets in R, closed sets which do not contain any intervals
(like N or any discrete closed set in R) etc. If a set (or its closure) contains an interval inside, then it can
never be nowhere-dense. Note that dense sets can not be nowhere-dense (so Q is not nowhere-dense in
A complete metric space can not be written as a denumerable union of nowhere-dense subsets.
of X for each n ∈ N. We can assume that Vn is closed and nowhere dense since ∪∞
i=1 Vn = X also for
such Vn . Now use the idea that Vn can not contain any balls inside, and so starting with any x1 , we can
get x2 ∈ B(x1 , r1 ) ∩ VnC giving B(x2 , r2 ) ⊂ B(x1 , r1 ) ∩ V1C , the latter being an open set. Now find such
a Cauchy sequence (xn ) from X, which can not converge in X, giving a contradiction.
Also, it is easy to see that the interior of a proper, closed subspace of any normed linear space is ϕ.
Theorem 3.3.3. Let X be any normed linear space and Y be a proper subspace of X. Then Y does not
Proof. Suppose Y contain an interior point, say y0 . Then we can show that Y will contain all points of
Corollary 3.3.4. Every finite-dimensional subspace of an infinite dimensional normed linear space is
nowhere dense.
Proof. Finite dimensional spaces are complete and hence closed. So by the above theorem, if Y is a finite-
0
dimensional subspace of an infinite dimensional space, then it is proper, closed and hence Y = ϕ.
Now we establish the connection between the existence of denumerable basis and completeness.
Theorem 3.3.5. Let X be a normed linear space with a denumerable basis, say {un }∞
n=1 . Then X can
X = ∪∞
n=1 Xn ,
where Xn = span{u1 , u2 , . . . , un }, finite dimensional space for each n ∈ N, and hence nowhere dense.
Example 3.3.6. Consider the spaces P[a, b] and c00 . These spaces have denumerable bases. Hence these
spaces are not Banach spaces w.r.to any norm. Further, from any normed space, collect a sequence of
linearly independent elements. Their span is a subspace of the normed linear space and has denumerable
basis. Hence such spaces can never be Banach spaces. Note that c00 is the span of (en ).
So in essence, if X has a finite basis, then it must be complete and if the basis is countable, then the
space can not be complete. Hence we need to tackle only the case when X has uncountable bases, to
3.4 Problems
1. Let X be a finite dimensional vectorspace with basis {u1 , u2 , . . . , un }. Show that for x ∈ X with
Pn
x= αi ui , αi ∈ K,
i=1
n
X
∥x∥∞ = ∥ αi ui ∥∞ = sup |αi |
i=1,2,...,n
i=1
is a norm on X.
n
!1/2
X
∥(x1 , x2 , . . . , xn )∥2 = |xi |2 where (x1 , x2 , . . . , xn ) ∈ Kn
i=1
and an arbitrary norm, say ∥ · ∥. Show directly (without using equivalence of norms on finite
dimensional normed linear spaces), that ∃ b > 0 such that ∥x∥ ≤ b∥x∥2 ∀x ∈ Kn .
3. Show that on a normed linear space X, ‘∥∥1 ≃ ∥∥2 if ∥∥1 is equivalent to ∥∥2 ’ gives an equivalence
5. Show that finite dimensional subspaces are closed in a normed linear space .
6. Every totally bounded sets in a metric space are bounded. Give example to show that bounded
sets may not be totally bounded. (Hint:- (R, d) with d(x, y) = min{|x − y|, 1})
9. Every compact sets in a metric space are totally bounded. Also show that converse is not true.
32
10. If A ⊂ X is totally bounded then all subsets of A are also totally bounded.
12. Suppose that there are two equivalent norms on a normed linear space. Then show that
• if a set is bounded w.r.to one norm, then it is bounded w.r.to the other norm also.
• if a set is closed w.r.to one norm, then it is closed w.r.to the other norm also.
• if a sequence is convergent w.r.to one norm, then it is convergent w.r.to the other norm also.
14. Show that the closed unit ball of ℓ2 is not compact, without using the theorem. (The same result
We have seen the connection between Banach spaces and the basis. Recall that a collection of vectors
β = {uα : α ∈ A} is a basis for a vector space (or a normed linear space) X if it is linearly independent
and it spans X. A basis always exists (by Zorn’s lemma) for any vector space and it may be finite,
denumerable or uncountable.
In this chapter, we plan to introduce some other basis-like concepts for a normed linear space . To
distinguish, we call the usual basis (using which every element of the space can be written as a finite
linear combination) a Hamel basis. Remember that we did not assume any condition on the norm of
the basis elements while defining the Hamel basis (as it is defined for vector spaces where we need not
have a norm in general). In a normed linear space, we can take the basis elements to be of norm 1;
and further, every element will have a unique finite linear combination expression in terms of the basis
vectors.
Since we have a topology on a normed linear space , we can think of using infinite linear combinations
to define a basis concept. This gives rise to a new concept of a basis for normed linear space .
Definition 4.2.1. Let X be a normed linear space and (un ) be a countable (finite or denumerable)
collection of unit elements from X. Such a collection is called a Schauder basis for X if every element
33
34
Theorem 4.2.3. If there is a countable Hamel basis for X, then that basis gives a Schauder basis (by
Proof. Take the countable basis, say (xn ) and form un = ∥xxnn ∥ , n ∈ N. Now being a Hamel basis every
XN
element in X has got a unique expression x = αn un , proving that it is a Schauder basis also.
n=1
Following is a simple example for a Schauder basis, which is not a Hamel basis.
Example 4.2.4. It is easy to see that (en ) is not a Hamel basis for ℓ1 , as by using a finite linear
But (en ) is a Schader basis for ℓ1 . Taking any x ∈ ℓ1 , we can write x = (x1 , x2 , . . .) and it is easy to
∞
X
show that x = xi ei , since for any ϵ > 0,
i=1
∞
X n
X ∞
X
∥x − sn ∥ = ∥ xi ei − xi ei ∥ = |xi | < ϵ
i=1 i=1 i=n+1
is possible for large n, being the tail of a convergent series, and the expression is unique.
Note that, for a finite-dimensional space, any Hamel basis consisting of unit vectors forms a Schauder
basis. It is also obvious that if there is a denumerable Hamel basis for a normed linear space , then
making the basis elements to be of norm 1 creates a Schauder basis. For example from the Hamel basis
There is a close connection between Schauder basis and separability of a normed linear space . We
Definition 4.2.5. A space X is said to be separable if there is a countable dense subset A in X. That
is A ⊆ X, A is countable and A = X.
Note that R is separable since rational numbers Q forms a countable dense subset. The space C is also
separable since Q + i Q is also countable and dense in C. Since a dense set intersects with all non-empty
Theorem 4.2.6. If there are uncountably many non-empty disjoint open sets in a topological space, then
Proof. If a dense set is there, it intersects with each of the open sets at different points, making it
Corollary 4.2.7. If a metric space has uncountably many points {xα }α∈A , where A is an uncountable
index set, which are at a distance of some ϵ0 > 0 or more, then the space can not be separable.
Proof. The collection of all {B(xα , ϵ0 ) : α ∈ A} forms an uncountable collection of disjoint open balls in
the space.
Proof. The collection {χ[a,a+ϵ) : 0 < ϵ < b − a} forms an uncountable set at 1 distance apart. Here χS
n
X
D={ ri ui : ri ∈ Q + i Q}.
i=1
We can observe that a Schauder basis is always linearly independent. However, they need not span
X (if they span, they form a Hamel basis itself). But the following is always true.
Now we establish the connection between Schauder basis and separable spaces.
Proof. Finite dimensional spaces are separable. Now, let (un ) be a Schauder basis for a normed linear
Xn
space X. Now form Dn = { rj uj : rj ∈ Q + i Q}, a countable set and let D = ∪∞
n=1 Dn , which is also
i=1
countable. Now show that D = X.
36
An example was constructed by Enflo in one of his renowned work to show that separable normed
linear space need not have a Schauder basis. We will see that for Hilbert spaces, it is true that it is
separable if and only if it has a countable orthonormal basis (which is also a Schauder basis).
The concept of separable spaces is very important in functional analysis as many results are proven
for these spaces. Also in Hilbert space theory, separable Hilbert spaces allow us to use a countable
Note that this concept is valid only in an inner product setting as it uses the inner product and orthogo-
nality concept. Recall that an orthonormal set is an orthogonal set of unit vectors. That is E = {uα }α∈A ,
Unlike the Hamel basis and Schauder basis, the orthonormal basis is not defined using the way elements
are represented using the basis elements (later we will see that such an expression is possible), but using
Definition 4.3.1. A maximal orthonormal set in an inner product space X is called an orthonormal
basis (or a complete orthonormal system) for X. That is E is an orthonormal basis if E is orthonormal
Do there exist orthonormal basis for every inner product space? The answer is ”yes”, if we assume
Zorn’s lemma.
Theorem 4.3.2. Let E0 be any orthonormal set in X. Then there exists an orthonormal basis for X
containing E0 .
Proof. Consider the set C of all orthonormal sets Eγ in X which contains E0 . This is a non empty set
with a partial order A ≺ B if A ⊆ B. Then every chain in this collection C has an upper bound, which
is their union. Hence there exists a maximal element, say E ∈ C. It is easy to show that this maximal
Note that orthonormal basis can be uncountable also (a Schauder basis is always countable).
Example 4.3.3. Consider ℓ2 and the collection E = {en }. Then it is easy to see that this collection
x ∈ Ẽ − E. But this x = (xn ), a sequence and xn = ⟨x, en ⟩ = 0 for all n ∈ N, giving a contradiction that
x = 0.
37
Same technique works for Kn with E = {e1 , e2 , . . . , en } (we can also use the fact that in an n−dimentional
space, there can’t be n + 1 orthonormal elements). It is very easy to see that every element in a fi-
nite dimensional inner product space can be expressed as a linear combination and the expression is
n
X
x= ⟨x, uj ⟩uj .
j=1
It is easy to verify the orthonormality of a given collection. The following result helps to check the
maximality.
Theorem 4.3.4. Let E be an orthonormal set in an inner product space X. Then E is an orthonormal
Using this proof, we can obtain the following result about Hamel basis.
Theorem 4.3.5. Suppose E is an orthonormal set and is also a basis for an inner product space X.
The following shows that given any orthonormal set in an inner product space X, it is an orthonormal
Proof. Let W = span(E) and V = span(E). Now E is a Hamel basis for W and so by above theorem,
For the second part, consider x ∈ span(E) such that x ∈ E ⊥ . Then there is a sequence xn ∈ span(E)
such that xn → x in span(E). Now x ∈ E ⊥ implies ⟨xn , x⟩ = 0 for all n ∈ N and hence ⟨xn , x⟩ → ⟨x, x⟩ =
0. Thus E ⊥ = {0}.
Remark 4.3.7. The above result helps us to conclude that {e1 , e2 , . . . , en } is an orthonormal basis for
Kn . Also (en ) is a basis for c00 . If we assume the fact that C00 is dense in ℓ2 , the same set (en ) is a basis
Now using Gramm-Schmidt process we can create a an orthonormal basis for (P[−1, 1], ∥∥2 ) from
the standard basis {1, x, x2 , . . .}. This set is actually the sequence of Legendre polynomials. If we assume
the fact that P[−1, 1] is dense in C[−1, 1], ∥∥2 , the same set of Legendre polynomials is an orthonormal
basis for C[−1, 1], ∥∥2 also, by means of part (ii) of the same. Now using the fact that C[−1, 1], ∥∥2 is
38
dense in L2 [−1, 1], we can see that Legendre polynomials is an orthonormal basis for L2 [−1, 1] by the
same result.
eint
Example 4.3.8. Consider un = √ for each n ∈ Z, which are in C[0, 2π]. It is easy to see that this set
2π
E = {un : n ∈ Z} is an orthonormal set . Hence it is an orthonormal basis for the span of E. The span
polynomials are dense (w.r.to ∥ · ∥∞ and so w.r.to ∥ · ∥2 also) in C[0, 2π]. Hence E is an orthonormal
basis for C[0, 2π] and since C[0, 2π] is dense in L2 [0, 2π], E is a countable orthonormal basis for L2 [0, 2π]
also.
Note that since eint = cos nt + i sin nt, the collection { √12π , sin
√ nt , cos
π
√ nt : n ∈ N} also forms an
π
orthonormal basis for the above three spaces using similar arguments.
Even though we called a maximal orthonormal set as an orthonormal basis , it is not clear in what
sense it is a basis. A basis’s fundamental role is to express a given element as a linear combination of the
basis elements. our next aim is to show this for orthonormal basis . But we need some standard results
to establish this fact. See problem 14, as we are expecting something like this in the general case also.
Let (un ) be a countable orthonormal set in an inner product space X and x ∈ X. Then
∞
X
|⟨x, uj ⟩|2 ≤ ∥x∥2 .
i=1
n
X n
X
Proof. Define sn = ⟨x, uj ⟩uj . Then ⟨x, sn ⟩ = ⟨sn , x⟩ = ⟨sn , sn ⟩ = |⟨x, uj ⟩|2 . Now
i=1 i=1
n
X
0 ≤ ∥x − sn ∥2 = ∥x∥2 − |⟨x, uj ⟩|2 ,
i=1
n
X
which implies that |⟨x, uj ⟩|2 ≤ ∥x∥2 for each n ∈ N, and taking n → ∞, we get the conclusion.
i=1
Now our aim is to see if, using an orthonormal basis E, can we have the representation
X
x= ⟨x, u⟩u
u∈E
39
for each x ∈ X?
Such a countable sum expression can’t be expected unless we have E to be a countable set. But if
except countably many ⟨x, u⟩ are zeros, then also we can have such an expression meaningful.
Theorem 4.3.11. Suppose E = {uα } is any orthonormal set in an inner product space X and x ∈ X.
{uα ∈ E : ⟨x, uα ⟩ =
̸ 0}
1
Proof. It is an application of Bessel’s inequality. First Observe that if we take Dn = {uα : ⟨x, uα ⟩ > n}
for n ∈ N, then D = ∪Dn . One can show that each Dn is a finite set, by showing that if it was not
m
so, by Bessel’s inequality we get ∥x∥2 ≥ n2 for any m ∈ N, a contradiction, since ∥x∥ must be a finite
number.
X
Thus for any x ∈ X, with respect to an orthonormal set E, the expression ⟨x, u⟩u is a countable
u∈E
expression, and hence there is a meaning for this sum. We can see that this expression turns out to
be the expression for x in a Hilbert space. In the following theorem in this regard, we need to use the
Let H be a Hilbert space and E be an orthonormal set in H Then E is an orthonormal basis if and
X
Proof. The converse part is simple. Assume x = ⟨x, uα ⟩uα . To show E is an orthonormal basis ,
uα ∈E
we need to show that E ⊥ = {0}. So take x ∈ E ⊥ . Then ⟨x, uα ⟩ = 0 for all uα , and hence x = 0.
X
For the other part, we will show that the series x = ⟨x, uα ⟩uα actually converge and call the limit
uα ∈E
as y, and then show that y = x.
Let H be a Hilbert space and E be an orthonormal set in H Then E is an orthonormal basis if and
Proof. It is easy to prove this result using the above result and by analysing the proof of Bessel’s inequality.
X X
We will show that |⟨x, uα ⟩|2 = ∥x∥2 if and only if x = ⟨x, uα ⟩uα . Note that both these
uα ∈E uα ∈E
summations are countable summations and the index set can be taken as N.
40
Xn
From Bessel’s inequality, we have ∥x − sn ∥2 = ∥x∥2 − |⟨x, ui ⟩|2 for each n ∈ N. Suppose that
X i=1
|⟨x, uα ⟩|2 = ∥x∥2 . This means that the RHS limit exists, and so the LHS limit also must exist,
uα ∈E
∞
X
which implies that x = lim sn = ⟨x, ui ⟩ui . Similar arguments give the converse part.
i=1
Now it looks like the orthonormal basis acts like a Schauder basis. Yes, it is. But by definition,
Proof. Note that in an orthonormal basis , ∥ui ∥ = 1 and by Fourier series representation, every element
has a unique representation in terms of the orthonormal basis elements. Note that in view of the Fourier
Hence by Theorem, every Hilbert space which has a countable basis is separable space. What about
the converse? The converse is also true in the Hilbert space settings. To show this, we first prove that
Theorem 4.3.15. . Suppose X is an inner product space and E0 be any orthonormal set in X. Then
partially ordered set with partial order ⊂ and every chain in E has an upper bound (the union). Hence
Now we prove the relation between separable Hilbert spaces and orthonormal basis .
Theorem 4.3.16. A Hilbert space is separable if and only if it has a countable orthonormal basis .
Proof. It there is a countable orthonormal basis , that is a Schauder basis also, and hence by Theorem,
each α ̸= β. Hence there are uncountably many disjoint open sets in H and hence it can not be separable,
4.4 Problems
2. If (X, ∥ · ∥∞ ) is a finite-dimensional normed linear space, then every Hamel basis of unit vectors is
5. Let a = (1, 1, · · · ). Show that {a, e1 , e2 , · · · } is a Schauder basis for the subspace c of ℓ∞ .
⊥
13. Prove that A⊥ = A .
14. Show that every inner product space has an orthonormal basis.
15. Show that an orthonormal basis is a Hamel basis if and only if the space is finite dimensional inner
converges.
42
Chapter 5
We want to introduce the continuity concept for linear transformations. This is possible since we consider
the domain and codomain to be normed linear space as they are topological spaces. Recall
T (αx + y) = αT x + T y ∀x, y ∈ X, α ∈ K.
Note that we consider only linear operators in this course as operators. The continuity of such
any ϵ > 0, there exists a δ > 0 such that ∥T x − T x0 ∥ < ϵ for all x ∈ X with ∥x − x0 ∥ < δ.
Since X and Y are metric spaces, the above definition is equivalent to saying that whenever any
sequence xn → x0 in X, T xn → T x0 as n → ∞.
Definition 5.1.1. bounded sets in normed spaces, unit ball closed and open, unit sphere.
Proof. see
Proof. see
43
44
Note that the above theorem says that T is bounded if and only if {T x : ∥x∥ ≤ 1} is a bounded set.
Theorem 5.1.4. T : X → Y is bounded if and only if there exists and α > 0 such that
∥T x∥ ≤ α∥x∥ ∀x ∈ X.
Proof. Write
Example 5.1.5. Consider a bounded sequence (λn ). Define a map T : ℓp → ℓp , 0 ≤ p < ∞, defined by
Then T is a bounded linear operator and α can be chosen as any upper bound of |λn |.
Example 5.1.6. Consider a function ϕ ∈ C[a, b]. Define a map T : (C[a, b], ∥ · ∥∞ ) → (C[a, b], ∥ · ∥∞ ),
defined by
Then T is a well defined bounded linear operator and α can be chosen as ∥ϕ∥∞ .
Then T is a well defined linear operator; but T is unbounded since the bounded set {xn : n ∈ N} is
1. T is bounded
2. T is continuous at 0
3. T is uniformly continuous on X
Proof. See
45
Proof. See
Proof. See
Proof. see
Proof. SEe
Theorem 5.3.3. If X is a finite dimensional normed linear space , then any operator T : X → Y is
bounded.
This means L(X, Y ) = B(X, Y ) if X is finite dimensional. The following result shows that it happens
5.4 Problems
2. Show that T : X → Y is bounded if and only if T (B(a, r)) is a bounded set for some a ∈ X and
r > 0.
Show that T is a bounded operator with respect to ∥ · ∥1 and not a bounded operator with respect
to ∥ · ∥2 .
∞
X x(j)
T (x1 , x2 , . . .) = ∀ (x1 , x2 , . . .) ∈ c00 .
j=1
j
Show that T is a bounded operator with respect to ∥ · ∥2 and not a bounded operator with respect
to ∥ · ∥∞ .
Contents
2.5 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
47
48
4.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
5.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45