Lecture Note Functional Analysis Chapter 2 Banach Spaces
Lecture Note Functional Analysis Chapter 2 Banach Spaces
1. Banach spaces
For p = ∞, we set
1
Example 2.1.4 (The space of continuous functions). Let Ω be a topo-
logical space, a subset of Rd , and let C(Ω) be the set of all continuous func-
tions f : Ω −→ C. The uniform/supremum norm on C(Ω) is given by
∥f ∥ = sup|f (x)|.
x∈Ω
Then (C(Ω), ∥·∥) is a Banach space. This is clearly that C(Ω) is a vector
space and the norm is easy to verify. Therefore, it suffices to show that C(Ω)
is complete. Let fn be a Cauchy sequence defined on ∥·∥, then
Hence the limit of fn exists and f (x) = limn→∞ fn (x). Fix some point x0 ∈ Ω,
since fn (x) is continuous at x0 , for every ϵ > 0, there exist δ > 0 such that
if 0 < |x − x0 | < δ, then
2
complete w.r.t. the supremum norm ∥·∥. The space C0 (Rd ) is the space of
all continuous functions vanishing at ∞. The latter means:
lim |f (x)| = 0
|x|→∞
which also equivalent to say that, for every ϵ > 0, there is a compact set
K such that |f (x)| < ϵ for all x ̸∈ K. The space (C0 (Rd ), ∥·∥) is a Banach
space. In total,
Exercise 2.1.5. Why CK (Rd ) is not complete, but C0 (Rd ) is complete w.r.t.
the supremum norm ∥·∥?
Solutions:
that converges to
sin x
x ̸= 0,
f (x) = x
1
x=0
3
as k → ∞. Then limx→±∞ f (x) = 0. We see that fk (x) ∈ CK (R) but
f ̸∈ CK (R), therefore CK (R) is not complete. Note that f ∈ C0 (R).
For given ϵ > 0, to prove the set
|f (x)| = |f (x) − fn (x) + fn (x)| ≤ |f (x) − fn (x)| + |fn (x)| < ϵ/2 + ϵ/2 = ϵ
For p = ∞, we define
4
Exercise 2.1.7. Write down the following inequality/facts from Lp -spaces:
(1). Minkowski’s inequality;
(2). Hölder’s inequality;
(3). Riesz-Fisher theorem.
Solutions:
Solutions:
5
Space of Bounded Operators
Definition 2.1.10.
(a). Let X, Y be two normed spaces with norm ∥·∥X , ∥·∥Y . A linear map
T : X −→ Y is bounded if there exists C ≥ 0 such that
∥T x∥Y ≤ C ∥x∥X .
(c). We denote L(X, Y ) or B(X, Y ) the space of all bounded operators from
X −→ Y . If X = Y , we write L(X) or B(X).
In fact, L(X, Y ) is a Banach space, with the norm ∥·∥op . In fact, X needs
not to be complete.
Proof. The fact that L(X, Y ) is a vector space. To prove ∥·∥op is a norm,
without loss of generality we set ∥x∥X = 1. We see that
= ∥T ∥op + ∥S∥op .
6
To prove completeness, let {Tn } be a Cauchy sequence in ∥·∥op . Given x ∈ X,
to show that {Tn x} is Cauchy, set ∥x∥X = 1, then
This shows that {Tn x} is Cauchy in Y . Since Y is Banach, then the limit
of Tn x is in Y , so T x = limn→∞ Tn x. To prove T x is in L(X, Y ), we need
to show that T x is linear and bounded. Obviously T x is linear. To prove
bounded, we see that
∥(Tn − T )x∥Y = lim ∥(Tn − Tm )x∥Y ≤ lim ∥(Tn − Tm )∥op ∥x∥X ≤ ϵ ∥x∥X .
m→∞ m→∞
7
Exercise 2.1.12. Show that if T, S ∈ L(X), then T S ∈ L(X). Moreover,
∥T S∥op ≤ ∥T ∥op ∥S∥op .
Proof. Note that L(X) is the space of all bounded operators from X to X.
If T, S ∈ L(X), then
Definition 2.1.13. Let ∥·∥1 , ∥·∥2 be two norms defined on a vector space X.
(a). ∥·∥1 is stronger than ∥·∥2 if there exists a M ≥ 1 such that for all x ∈ X,
∥x∥2 ≤ M ∥x∥1 .
Remark. The definition 2.14 (a) asserts that the convergence sequence in
norm 1 is less than norm 2.
8
Exercise 2.1.14. Denote ∥·∥1 ≺ ∥·∥2 if latter is strong than the former, and
∥·∥1 ∼ ∥·∥2 if the two norms are equivalent.
(a). Show that ≺ is a partial order.
(b). Show that ∼ is an equivalence relation.
9
This implies that M −1 N −1 ∥x∥1 ≤ N −1 ∥x∥2 ≤ ∥x∥3 ≤ N ∥x∥2 ≤ M N ∥x∥1 .
This shows that ∥x∥1 ∼ ∥x∥3 . Hence ∼ is an equivalence relation.
Exercise 2.1.16. Show that if two norms ∥·∥1 , ∥·∥2 over the space X are
equivalent, then (X, ∥·∥1 ), (X, ∥·∥2 ) are isomorphic.
which is clearly linear and bijective. Since ∥·∥1 and ∥·∥2 are equivalent, then
by definition
10
Example 2.1.17. Let ∥·∥p be the p-norm over Cd , with d < ∞. Then ∥·∥p
are equivalent with each other for all p ∈ [1, ∞]. Particularly, given x ∈ Cd ,
we have
d
X 1/p
p
∥x∥∞ ≤ ∥x∥p ≤ d ∥x∥∞ ⇐⇒ max|xk | ≤ |xk | ≤ d max|xk |.
k k
k=1
Hence, ∥·∥p ∼ ∥·∥∞ . The equivalence relation over ∼ shows every norm are
equivalent. However, the space (Cd , ∥·∥p ) are not isometry to any different p.
See Exercise below. Note that also the bound becomes worse when d → ∞.
Exercise 2.1.18. Show that (R2 , ∥·∥1 ) are not isometric to (R2 , ∥·∥2 ). Par-
ticularly, the unit ball of ∥·∥1 is a square, and that of ∥·∥2 is a unit disk. It
is possible to bijectively map a square to a disk via linear transform?
which is clearly they are not equal. Hence T is not isometry between (R2 , ∥·∥1 )
and (R2 , ∥·∥2 ). It is not possible.
Exercise 2.1.19.
(a). If (Ω, µ) is bounded measure space, that is µ(Ω) = 1, which norm is
stronger? ∥·∥L1 or ∥·∥L2 .
(b). Show that the two norms on L1 (Rd ) and L2 (Rd ) are not equivalent.
11
Then, since f is integrable, then f is bounded, so |f (x)| ≤ M . Hence,
Z Z
2
∥f ∥L2 = |f (x)| dµ(x) ≤ M |f (x)|dµ(x) = M ∥f ∥L1 .
Ω Ω
The right inequality holds, but the left inequality does not hold, a contradic-
tion.
Definition 2.2.1. Let X be a Banach space. The dual space X ∗ is the vector
space of all bounded linear functionals ℓ : X −→ C. Boundedness is defined
as: there exists M ≥ 0 such that
|ℓ(x)| ≤ M ∥x∥X , x ∈ X.
The norm of ℓ is defined as the smallest M > 0 such that the above holds,
that is, the operator norm of ℓ:
|ℓ(x)|
∥ℓ∥X ∗ = sup = sup |ℓ(x)|.
x:x̸=0 ∥x∥ x:∥x∥=1
12
A hint to generalize the inner product on Hilbert spaces to Banach is Riesz
representation theorem, which says inner product x 7→ (v, x) is in fact a
linear functional ℓv , and vice versa. That is we may view (·, ·) : H × H −→ C
as the map ⟨·, ·⟩ : H∗ × H −→ C, where
That is, taking the inner product between v, x is equivalent to test x against
a linear functional ℓv ∈ H∗ . Hence, to define inner product for a Banach
space X, we have
⟨·, ·⟩ : X ∗ × X −→ C,
where the bilinear map ⟨·, ·⟩ takes a linear functional ℓ over X and a member
x ∈ X to produce a complex number ⟨ℓ, x⟩ = ℓ(x) ∈ C.
Theorem 2.2.4. Let p ∈ [1, ∞). Then the dual space of Lp (Ω, µ) (for some
measure space (Ω, µ)) is isomorphic to Lq (X, µ), where p, q are Hölder dual:
p−1 + q −1 = 1.
13
Clearly, it defines a linear functional. Moreover, by Hölder’s inequality,
where ϕ(x) is the phase angle such that g(x)f (x) is real. One have to verify
also g ∈ Lp (using p−1 + q −1 = 1).
Surjective of T is technical. One needs to show every linear functional ℓ
arisen as ℓg for some g. We may seek help from more powerful representation
theorem, such has Riesz representation theorem for L2 , or an even powerful
one, Radon-Nykodym theorem.
Example 2.2.5. The dual space of sequence spaces can be easily identified,
because sequence spaces are essentially Rd for d = ∞. Some techniques from
Rd may apply in this case. For instance,
(i). (ℓp )∗ ∼
= ℓq for p ∈ [1, ∞) and p−1 + q −1 = 1;
(ii). (ℓ1 )∗ ∼
= ℓ∞ as above. We also have c∗0 ∼= ℓ1 , where c0 ⊂ ℓ∞ is the space
of sequences that vanishes as the term → ∞.
14
Definition 2.2.6 (Double dual). The double dual X ∗∗ of a Banach space
is the dual space of its dual space X ∗ , that is, X ∗∗ = (X ∗ )∗ .
The double dual space X ∗∗ of X is larger than X, see theorem below. But
in the case when X ∗∗ ∼ = X, we say the space X is reflexive. For instance, a
= H∗ ∼
Hilbert space is reflexive since H∗∗ ∼ = H, and so are Lp for p ∈ (1, ∞)
= (Lq )∗ ∼
since (Lp )∗∗ ∼ = Lp . The space L1 (Ω), C0 (Ω) are not.
Proof. For each x ∈ X, let ℓ̂(·) be the linear functional on X ∗ which assigns
to each λ ∈ X ∗ the number λ(x). Since
ℓ̂ = ∥x∥X
X ∗∗
15
3. Hahn-Banach theorem
Let us work with a simple example to show how Hann-Banach can be useful.
Consider a Hilbert space H and let S ⊂ H be a subspace. Our goal here is
to find an annihilator y ∈ H, y ̸= 0 such that
|ℓ(x)| ≤ p(x), x ∈ S,
16
Proof. The finite dimension X is easy, given S ⊂ X, and ℓ pre-defines on
S, we add a vector x1 ∈ X \ S, and we extend ℓ to S1 = span(S ∪ {x1 }) so
that the extension ℓ1 still satisfies the bound |ℓ1 (x)| ≤ p(x) for all x in the
extended subspace. Then we add another vector x2 ∈ X \ S1 . We continuous
this until the vector space is exhausted. The final extension will be what we
need.
This argument cannot be worked out for the case of infinite dimensions. We
need some argument which says if things can be done for finite case, it can
also be done on the infinite case. Therefore, we need Zorn’s lemma.
We first prove the case of real Banach spaces, then the complex case follows
by complexification.
Step 1: We show we can always extend ℓ : S −→ R from S ̸= X to ℓ : S ′ −→
R, where S ′ has one more (dimension) than S. As mentioned earlier, we take
v ∈ X \ S, and we define S ′ = span(S ∪ {v}) and ℓ : S ′ −→ R by
17
Rearranging terms: for all x, y ∈ S,
Step 2: Consider
18
ℓ is a desired extension of ℓ.
Proof. Let Y be the subspace consisting of all scalar multiples of y and define
λ(ay) = a ∥y∥. By using Corollary 2.3.2, we can construct Λ with ∥Λ∥ = ∥λ∥
extending λ to all of X. But, since Λ(y) = ∥y∥, ∥Λ∥ = 1 and therefore
19
4. Axiom of Choices and Baire Category Theorem
The axiom of choice seems to be very intuitive. In fact, there are quite
some counter intuitive (often wild) consequence follows.
1. Zorn’s lemma-it is proved Zorn’s lemma is equivalent to AC;
2. Well-ordering theorem-every set can be well-ordered. It is also proved this
is equivalent to AC.
3. Banach-Tarski theorem-one can construct two balls of the same weight
from one using piecewise rotation and translation;
4. Existence of non-measurable set;
5. Tarski’s theorem-A has the same cardinality as A × A.
20
Theorem 2.5.3 (Baire’s category). A complete metric space X is not of
first category.
S
Proof. If X is of first category, we write X = k Fk , where Fk ’s are nowhere
dense. We assume that Fk are closed (Note: If Fk is nowhere dense, so
is its closure). Construct a sequence {xk }k as follows: Take x1 such that
Br1 (x1 ) ∈ X \ F1 (because X \ F1 is open). Take x2 ∈ Br1 (x1 ) \ F2 . The set
is nonempty because F2 is nowhere dense (otherwise Br1 (x1 ) ⊂ F2 , and F2
cannot be nowhere dense). Since Br1 (x1 )\F2 is open, we have some r2 ≤ r1 /2
such that Br (x2 ) ⊂ Br1 (x1 ) \ F2 . Repeat the construction we get a sequence
S
{xk }k with the property that xk ̸∈ j≤k Fj . The sequence is Cauchy from the
construction (because {xk }k≥n ∈ Br/2n (xn )). The completeness leads to the
existence of some element x ∈ X which is not in any Fk ’s, a contradiction.
Proof. If T is bounded (that is, also continuous by Problem 1.4), then the
preimage of B1Y (0) is open, and hence T −1 (B1Y (0)) has an interior point. Con-
versely, if T −1 (B1Y (0)) has an interior point, namely, Bϵ (x0 ) ⊂ T −1 (B1Y (0)).
21
Pick x ∈ Bϵ (0), then
The right hand side does not depend on x. Scaling x by ϵ−1 we find for all
x ∈ B1X (0) we have
∥T x∥Y ≤ (1 + C)/ϵ.
Hence T is bounded.
Proof. Fill up the space X with a collection of closed set {Bn } with
\
Bn = {x ∈ X : ∥Tα x∥Y ≤ n, ∀α} = Tα−1 (BnY (0)).
α
S
We need to show that X = n Bn . Since X is complete, a consequence of
Baire category theorem shows one of these sets, say Bn is not nowhere dense
0
for some n, that is, Bn0 = Bn ̸= ∅. That is, Bn contains an open ball. Now
we proceed with the same argument from the proof of the theorem 2.5.4.
First we have Bϵ (x0 ) ⊂ Bn . Then for all x ∈ Bϵ (0),
22
where the right hand side is independent of x ∈ Bϵ (0). Scaling x by ϵ−1 we
find
= T (Bn − x)
= T (Bn (−x))
⊂ T (BR (0))
23
If we prove
T (B1 ) ⊂ T (B2 ),
then this proves T (B2 ) has an interior point. By scaling and translation, this
shows that T (Bϵ (x)) has an interior point. This then shows that T (U ) is
open.
It remains to prove the inclusion. Let ϵ > 0 be from above. Pick y ∈ T (B1 ),
and ∥y − T (x1 )∥Y < ϵ/2 (by the definition of closure), that is,
Y
y − T x1 ∈ Bϵ/2 ⊂ T (B1/2 ).
k=1
∗
P∞
Define x = k=1 xk . The series is summable because ∥xk ∥ ≤ 2−k . Moreover,
x∗ ∈ B2 . Passing n → ∞, we find
T (x∗ ) = y ∈ T (B2 ).
24
Proof. To show T −1 is bounded, it suffices to show that T −1 is continuous.
For any open set U ⊂ X, consider T (U ) in Y . Since T is bijective bounded
linear map, then by open mapping theorem T (U ) is open, so T −1 is contin-
uous.
π1 : (x, T x) −→ x, π2 : (x, T x) −→ T x.
25
Corollary 2.5.10 (Helinger-Toplitz). Let A : H −→ H be a linear oper-
ator, where H is a Hilbert space. Suppose A satisfies
Then A is bounded.
Proof. We will prove that Γ(A) is closed. Suppose that (xn , Axn ) → (x, y).
We need only prove that (x, y) ∈ Γ(A), that is, that y = Ax. But, for any
z ∈ H,
26