Fumctional Analysis 2024-25
Fumctional Analysis 2024-25
CHI-WAI LEUNG
Abstract. This is a note on the course MATH 4010: Functional analysis in 2024-25, 1st term.
1. Normed spaces
Definition 1.1. Let X be a vector space over a field K, where K = R or C. A function ∥·∥ : X → R
is called a norm on X if it satisfies the following conditions.
(i) ∥x∥ ≥ 0 for all x ∈ X.
(ii) ∥x∥ = 0 if and only if x = 0.
(iii) ∥αx∥ = |α|∥x∥ for x ∈ X and α ∈ K.
(iii) (Triangle inequality) ∥x − y∥ ≤ ∥x − z∥ + ∥z − y∥ for all x, y, z ∈ X.
In this case, the pair (X, ∥ · ∥) is called a normed space.
Example 1.2. The following are important examples of finite dimensional normed spaces.
(n)
(i) Let ℓ∞ = {(x1 , ..., xn ) : xi ∈ K, i = 1, 2..., n}. Put ∥(x1 , ..., xn )∥∞ = max{|xi | : i = 1, .., n}.
(n)
(ii) Let ℓp = {(x1 , ...., xn ) : xi ∈ K, i = 1, 2..., n}. Put ∥(x1 , ..., xn )∥p = ( ni=1 |xi |p )1/p for
P
1 ≤ p < ∞.
Proposition 1.3. If X is a normed space, then the addition (x, y) ∈ X × X 7→ x + y ∈ X and the
scalar multiplication (α, x) ∈ K × X 7→ αx ∈ X both are continuous maps.
Notation 1.4. From now on, (X, ∥ · ∥) always denotes a normed space over a field K.
For r > 0 and x ∈ X, let
(i) B(x, r) := {y ∈ X : ∥x − y∥ < r} (called an open ball with the center at x of radius r) and
B ∗ (x, r) := {y ∈ X : 0 < ∥x − y∥ < r}
(ii) B(x, r) := {y ∈ X : ∥x − y∥ ≤ r} (called a closed ball with the center at x of radius r).
Put BX := {x ∈ X : ∥x∥ ≤ 1} and SX := {x ∈ X : ∥x∥ = 1} the closed unit ball and the unit
sphere of X respectively.
Definition 1.5. We say that a sequence (xn ) in X converges to an element a ∈ X if lim ∥xn −a∥ =
0, i.e., for any ε > 0, there is N ∈ N such that ∥xn − a∥ < ε for all n ≥ N .
In this case, (xn ) is said to be convergent and a is called a limit of the sequence (xn ).
Remark 1.7. Using the notations as above, a point z ∈ A if and only if B(z, r) ∩ A ̸= ∅ for all
r > 0. This is equivalent to saying that there is a sequence (xn ) in A such that xn → a. In fact,
this can be shown by considering r = n1 for n = 1, 2....
Proposition 1.8. Using the notations as before, we have the following assertions.
(i) A is closed in X if and only if its complement X \ A is open in X.
(ii) The closure A is the smallest closed subset of X containing A. The ”smallest” in here
means that if F is a closed subset containing A, then A ⊆ F .
Consequently, A is closed if and only if A = A.
Proof. If A is empty, then the assertions (i) and (ii) both are obvious. Now assume that A ̸= ∅.
For part (i), let C = X \ A and b ∈ C. Suppose that A is closed in X. If there exists an element
b ∈ C \ int(C), then B(b, r) ⫅̸ C for all r > 0. This implies that B(b, r) ∩ A ̸= ∅ for all r > 0 and
hence, b is a limit point of A since b ∈ / A. It contradicts to the closeness of A. Thus, C = int(C)
and thus, C is open.
For the converse of (i), assume that C is open in X. Assume that A has a limit point z but z ∈ / A.
Since z ∈ / A, z ∈ C = int(C) because C is open. Hence, we can find r > 0 such that B(z, r) ⊆ C.
This gives B(z, r) ∩ A = ∅. This contradicts to the assumption of z being a limit point of A. Thus,
A must contain all of its limit points and hence, it is closed.
For part (ii), we first claim that A is closed. Let z be a limit point of A. Let r > 0. Then there
is w ∈ B ∗ (z, r) ∩ A. Choose 0 < r1 < r small enough such that B(w, r1 ) ⊆ B ∗ (z, r). Since w is a
limit point of A, we have ∅ = ̸ B ∗ (w, r1 ) ∩ A ⊆ B ∗ (z, r) ∩ A. Hence, z is a limit point of A. Thus,
z ∈ A as required. This implies that A is closed.
It is clear that A is the smallest closed set containing A.
The last assertion follows from the minimality of the closed sets containing A immediately.
The proof is complete. □
A sequence (xn ) in X is called a Cauchy sequence if for any ε > 0, there is N ∈ N such that
∥xm − xn ∥ < ε for all m, n ≥ N . We have the following simple observation.
Proposition 1.11. Let X be a normed space. Then the following assertions are equivalent.
(i) X is a Banach
P∞ space. P∞
P∞ n=1 xn is absolutely convergent in X, i.e., n=1 ∥xn ∥ < ∞, implies that the
(ii) If a series
series n=1 xn converges in the norm.
Proof. (i) ⇒ (ii) is obvious.
Now suppose that Part (ii) holds. Let (yn ) be a Cauchy sequence in X. It suffices to show that
(yn ) has a convergent subsequence. In fact, by the definition of a Cauchy sequence, there is a
subsequence (ynk ) such that ∥ynk+1 − ynk ∥ < 21k for all k = 1, 2.... By the assumption, the series
P∞
k=1 (ynk+1 − ynk ) converges in the norm, and hence the sequence (ynk ) is convergent in X. The
proof is complete. □
3
Example 1.14. Let X be a locally compact Hausdorff space, for example, K. Let C0 (X) be the
space of all continuous K-valued functions f on X which are vanish at infinity, i.e., for every ε > 0,
there is a compact subset D of X such that |f (x)| < ε for all x ∈ X \ D. Now C0 (X) is endowed
with the sup-norm, i.e.,
∥f ∥∞ = sup |f (x)|
x∈X
for every f ∈ C0 (X). Then C0 (X) is a Banach space. (Try to prove this fact for the case
X = R. Just use the knowledge from MATH 2060 !!!)
Proposition 1.15. Let (X, ∥ · ∥) be a normed space. Then there is a normed space (X0 , ∥ · ∥0 ),
together with a linear map i : X → X0 , satisfies the following conditions.
(i) X0 is a Banach space.
(ii) The map i is an isometry, that is, ∥i(x)∥0 = ∥x∥ for all x ∈ X.
(iii) the image i(X) is dense in X0 , that is, i(X) = X0 .
Moreover, such pair (X0 , i) is unique up to isometric isomorphism in the following sense.
If (W, ∥ · ∥1 ) is a Banach space and an isometry j : X → W is an isometry such that j(X) = W ,
then there is an isometric isomorphism ψ from X0 onto W such that
j = ψ ◦ i : X → X0 → W.
4 CHI-WAI LEUNG
Example 1.16. Proposition 1.15 cannot give an explicit form of the completion of a given normed
space. The following examples are basically due to the uniqueness of the completion.
(i) If X is a Banach space, then the completion of X is itself.
(ii) The completion of the finite sequence space c00 is the null sequence space c0 .
(iii) The completion of Cc (R) is C0 (R).
Definition 2.1. Two norms ∥·∥ and ∥·∥′ on a vector space X are equivalent, denoted by ∥·∥ ∼ ∥·∥′ ,
if there are positive numbers c1 and c2 such that c1 ∥ · ∥ ≤ ∥ · ∥′ ≤ c2 ∥ · ∥ on X.
Example 2.2. Consider the norms ∥·∥1 and ∥·∥∞ on ℓ1 . We want to show that ∥·∥1 and ∥·∥∞ are
not equivalent. In fact, if we put xn (i) := (1, 1/2, ..., 1/n, 0, 0, ....) for n, i = 1, 2.... Then xn ∈ ℓ1
for all n. Note that (xn ) is a Cauchy sequence with respect to the norm ∥ · ∥∞ but it is not a Cauchy
sequence with respect to the norm ∥ · ∥1 . Hence ∥ · ∥1 ≁ ∥ · ∥∞ on ℓ1 .
Proposition 2.3. All norms on a finite dimensional vector space are equivalent.
Proof. LetPX be a finite dimensional vector space and let {e1 , ..., en } be a vector basis of X. For
each x = ni=1 αi ei for αi ∈ K, define ∥x∥0 = maxni=1 |αi |. Then ∥ · ∥0 is a norm X. The result is
obtained by showing that all norms ∥ · ∥ on X are equivalent
X to ∥ · ∥0 .
Pn
Note that for each x = i=1 αi ei ∈ X, we have ∥x∥ ≤ ( ∥ei ∥)∥x∥0 . It remains to find c > 0
1≤i≤n
such that c∥ · ∥0 ≤ ∥ · ∥. In fact, let SX := {x ∈ X : ∥x∥0 = 1} be the unit sphere of X with respect
to the norm ∥ · ∥0 . Note that by using the Weierstrass Theorem on K, we see that SX is compact
with respect to the norm ∥ · ∥0 .
Define a real-valued function f on the unit sphere SX of X by
f : x ∈ SX 7→ ∥x∥.
Note that f > 0 and f is continuous with respect to the norm ∥ · ∥0 . Hence, there is c > 0 such
that f (x) ≥ c > 0 for all x ∈ SX . This gives ∥x∥ ≥ c∥x∥0 for all x ∈ X as desired. The proof is
complete. □
In the remainder of this section, we want to show that the converse of Corollary 2.4(ii) holds.
Before this result, we need the following useful result.
Lemma 2.5. Riesz’s Lemma: Let Y be a closed proper subspace of a normed space X. Then for
each θ ∈ (0, 1), there is an element x0 ∈ SX such that d(x0 , Y ) := inf{∥x0 − y∥ : y ∈ Y } ≥ θ.
Proof. Let u ∈ X − Y and d := inf{∥u − y∥ : y ∈ Y }. Note that since Y is closed, d > 0 and
hence we have 0 < d < dθ because 0 < θ < 1. This implies that there is y0 ∈ Y such that
u−y0
0 < d ≤ ∥u − y0 ∥ < dθ . Now put x0 := ∥u−y 0∥
∈ SX . We are going to show that x0 is as desired.
Indeed, let y ∈ Y . Since y0 + ∥u − y0 ∥y ∈ Y , we have
1
∥x0 − y∥ = ∥u − (y0 + ∥u − y0 ∥y)∥ ≥ d/∥u − y0 ∥ > θ.
∥u − y0 ∥
Thus, d(x0 , Y ) ≥ θ. □
Remark 2.6. The Riesz’s lemma does not hold when θ = 1. The following example can be found
in the Diestel’s interesting book without proof (see [6, Chapter 1 Ex.3(i)]).
R1
Let X = {x ∈ C([0, 1], R) : x(0) = 0} and Y = {y ∈ X : 0 y(t)dt = 0}. Both X and Y are
endowed with the sup-norm. Note that Y is a closed proper subspace of X. We are going to show
that for any x ∈ SX , there is y ∈ Y such that ∥x − y∥∞ < 1. Thus, the Riesz’s Lemma does not
hold as θ = 1 in this case.
In fact, let x ∈ SX . Since x(0) = 0 with ∥x∥∞ = 1, we can find 0 < a < 1/4 such that |x(t)| ≤ 1/4
for all t ∈ [0, a].
We fix 0 < ε < 1/4 first. Since x is uniform continuous on [a, 1], we can find a partitions a = t0 <
· · · < tn = 1 on [a, 1] such that sup{|x(t) − x(t′ )| : t, t′ ∈ [tk−1 , tk ]} < ε/4. Now for each (tk−1 , tk ),
if sup{x(t) : t ∈ [tk−1 , tk ]} > ε, then we set ϕ(t) = ε. In addition, if inf{x(t) : t ∈ [tk−1 , tk ]} < −ε,
then we set ϕ(t) = −ε. From this, one can construct a continuous function ϕ on [a, 1] such that
R1
∥ϕ − x|[a,1] ∥∞ < 1 and |ϕ(x)| < 2ε for all x ∈ [a, 1]. Hence, we have | a ϕ(t)dt| ≤ 2ε(1 − a).
As |x(t)| < 1/4 on [0, a], so if we choose ε small enough such that (1 − a)(2ε) < a/4, then we can
find a continuous function y1 on [0, a] such that |y1 (t)| < 1/4 on [0, a] with y1 (0) = 0; y1 (a) = x(a)
Ra R1
and 0 y1 (t)dt = − a ϕ(t)dt. Now we define y = y1 on [0, a] and y = ϕ on [a, 1]. Then ∥y −x∥∞ < 1
and y ∈ Y is as desired.
Theorem 2.7. X is a finite dimensional normed space if and only if the closed unit ball BX of X
is compact.
Proof. The necessary condition has been shown by Proposition 2.4(ii).
Now assume that X is of infinite dimension. Fix an element x1 ∈ SX . Let Y1 = Kx1 . Then
Y1 is a proper closed subspace of X. The Riesz’s lemma gives an element x2 ∈ SX such that
∥x1 − x2 ∥ ≥ 1/2. Now consider Y2 = span{x1 , x2 }. Then Y2 is a proper closed subspace of X since
dim X = ∞. To apply the Riesz’s Lemma again, there is x3 ∈ SX such that ∥x3 − xk ∥ ≥ 1/2 for
k = 1, 2. To repeat the same step, there is a sequence (xn ) ∈ SX such that ∥xm − xn ∥ ≥ 1/2 for
all n ̸= m. Thus, (xn ) is a bounded sequence without any convergence subsequence. Hence, BX is
not compact. The proof is complete. □
Recall that a metric space Z is said to be locally compact if for any point z ∈ Z, there is a
compact neighborhood of z. Theorem 2.7 implies the following corollary immediately.
Corollary 2.8. Let X be a normed space. Then X is locally compact if and only if dim X < ∞.
In view of the proof of Theorem 2.7, we call a subset A of a normed space X a r-discrete set if
∥x − y∥ > r whenever x, y ∈ A with x ̸= y.
6 CHI-WAI LEUNG
Proposition 2.9. Let X be an infinite dimensional normed space and r > 0. If A := {xi }i∈I is
an infinite bounded r-discrete subset of X, then |I| ≤ dim X. Consequently, if {Bi : i ∈ I} is a
collection of pairwise disjoint r/2-balls in a bounded subset of X, then |I| ≤ dim X.
Proof. Let C be the collection of linearly independent subsets of A. Then the Zorn’s Lemma gives
a maximal element {xi : i ∈ J} in C for a subset J of I. We are going to show that |J| = |I|.
Suppose not. We assume that |J| < |I|. Now for each finite subset F of J, put VF the linear span
of the set F . Since VF is of finite dimension, the set VF ∩ A has finite volume with respect to the
Lebesgue measure in VF . Then VF ∩ A is a finite set because it is a r-discrete set. Therefore, if we
let F(J) := {F ⊆ J : |F | < ∞}, then we have
[
| VF ∩ A| = |F(J)| = |J|.
F ∈F(J)
[
Hence, if |J| < |I|, then VF ∩ A ⊊ A. Hence, there is an element xi0 ∈ A so that xi0 ∈
/ VF ∩ A
F ∈F(J)
for all F ∈ F(J). This implies that xi0 is linearly independent to the set (xi )i∈J . This contradicts
to the maximality of (xi )i∈J . Hence, |I| = |J| ≤ dim X by the definition of dimension. □
Proposition 3.3. Let X and Y be normed spaces. Let B(X, Y ) be the set of all bounded linear
maps from X into Y . For each element T ∈ B(X, Y ), let
∥T ∥ = sup{∥T x∥ : x ∈ BX }.
be defined as in Proposition 3.1.
Then (B(X, Y ), ∥ · ∥) becomes a normed space.
Furthermore, if Y is a Banach space, then so is B(X, Y ).
In particular, if Y = K, then B(X, K) is a Banach space. In this case, put X ∗ := B(X, K) and call
it the dual space of X.
Proof. We can directly check that B(X, Y ) is a normed space (Do It By Yourself !).
We want to show that B(X, Y ) is complete if Y is a Banach space. Let (Tn ) be a Cauchy sequence in
B(X, Y ). Then for each x ∈ X, it is easy to see that (Tn x) is a Cauchy sequence in Y . Thus, lim Tn x
exists in Y for each x ∈ X because Y is complete. Hence, we can define a map T x := lim Tn x ∈ Y
for each x ∈ X. Clearly, T is a linear map from X into Y .
We need show that T ∈ B(X, Y ) and ∥T − Tn ∥ → 0 as n → ∞. Let ε > 0. Since (Tn ) is a Cauchy
sequence in B(X, Y ), there is a positive integer N such that ∥Tm − Tn ∥ < ε for all m, n ≥ N .
Hence, we have ∥(Tm − Tn )(x)∥ < ε for all x ∈ BX and m, n ≥ N . Taking m → ∞, we have
∥T x − Tn x∥ ≤ ε for all n ≥ N and x ∈ BX . Therefore, we have ∥T − Tn ∥ ≤ ε for all n ≥ N . From
this, we see that T − TN ∈ B(X, Y ) and thus, T = TN + (T − TN ) ∈ B(X, Y ) and ∥T − Tn ∥ → 0
as n → ∞. Therefore, limn Tn = T exists in B(X, Y ). □
Remark 3.4. By using Proposition 3.1, we can show that if f : X → K is any linear functional
defined on a vector space X, then X can be endowed with a norm so that f is bounded. P
In fact, if we fix a vector basis (ei )i∈I for X and put ∥x∥∞ := maxi∈I |ai | as x = i∈I ai ei ∈ X,
(note that it is a finite sum), where ai ∈ K, then the function ∥ · ∥∞ is a norm on X. Now for each
x ∈ X, set
∥x∥1 := |f (x)| + ∥x∥∞ .
Clearly, the function ∥ · ∥1 is a norm on X. In addition, we have |f (x)| ≤ ∥x∥1 for all x ∈ X.
Hence, f is bounded on X with respect to the norm ∥ · ∥1 as required.
Proposition 3.5. Let X and Y be normed spaces. Suppose that X is of finite dimension n. Then
we have the following assertions.
(i) Any linear operator from X into Y must be bounded.
(ii) If Tk : X → Y is a sequence of linear operators such that Tk x → 0 for all x ∈ X, then
∥Tk ∥ → 0.
Proof. Using Proposition 2.3 and the notations as in the proof, then there is c > 0 such that
Xn n
X
|αi | ≤ c∥ αi ei ∥
i=1 i=1
for all scalars α1 , ..., αn . Therefore, for any linear map T from X to Y , we have
∥T x∥ ≤ max ∥T ei ∥ c∥x∥
1≤i≤n
for all x ∈ X. This gives the assertions (i) and (ii) immediately. □
Remark 3.6. The assumption of X of finite dimension in Proposition 3.5 cannot be removed. For
example, if for each positive integer k, we define fk : c0 → R by fk (x) := x(k), then fk is bounded
for each k and
lim fk (x) = lim x(k) = 0
k→∞ k→∞
for all x ∈ c0 . However fk ↛ 0 because ∥fk ∥ ≡ 1 for every k.
8 CHI-WAI LEUNG
Proposition 3.7. Let Y be a closed subspace of X and X/Y be the quotient space. For each
element x ∈ X, put x̄ := x + Y ∈ X/Y the corresponding element in X/Y . Define
If we let π : X → X/Y be the natural projection, i.e., π(x) = x̄ for all x ∈ X, then (X/Y, ∥ · ∥)
is a normed space and π is bounded with ∥π∥ ≤ 1. In particular, ∥π∥ = 1 as Y is a proper closed
subspace.
Furthermore, if X is a Banach space, then so is X/Y .
In this case, we call ∥ · ∥ in (3.1) the quotient norm on X/Y .
Proof. Note that since Y is closed, we can directly check that ∥x̄∥ = 0 if and only is x ∈ Y , i.e.,
x̄ = 0̄ ∈ X/Y . It is easy to check the other conditions of the definition of a norm. Thus, X/Y is a
normed space. Moreover, π is clearly bounded with ∥π∥ ≤ 1 by the definition of the quotient norm
on X/Y .
Furthermore, if Y ⊊ X, then by using the Riesz’s Lemma 2.5, we see that ∥π∥ = 1.
We show the last assertion. Suppose that X is a Banach space. Let (x̄n ) be a Cauchy sequence in
X/Y . It suffices to show that (x̄n ) has a convergent subsequence in X/Y .
Indeed, since (x̄n ) is a Cauchy sequence, we can find a subsequence (x̄nk ) of (x̄n ) such that
for all k = 1, 2.... Then by the definition of quotient norm, there is an element y1 ∈ Y such that
∥xn2 − xn1 + y1 ∥ < 1/2. Note that we have, xn1 − y1 = x̄n1 in X/Y . Thus, there is y2 ∈ Y such
that ∥xn2 − y2 − (xn1 − y1 )∥ < 1/2 by the definition of quotient norm again. In addition, we have
xn2 − y2 = x̄n2 . Then we also have an element y3 ∈ Y such that ∥xn3 − y3 − (xn2 − y2 )∥ < 1/22 .
To repeat the same step, we can obtain a sequence (yk ) in Y such that
for all k = 1, 2.... Therefore, (xnk − yk ) is a Cauchy sequence in X and thus, limk (xnk − yk ) exists
in X while X is a Banach space. Set x = limk (xnk − yk ). On the other hand, note that we have
π(xnk − yk ) = π(xnk ) for all k = 1, 2, , ,. This tells us that limk π(xnk ) = limk π(xnk − yk ) = π(x) ∈
X/Y since π is bounded. Therefore, (x̄nk ) is a convergent subsequence of (x̄n ) in X/Y . The proof
is complete. □
Corollary 3.8. Let T : X → Y be a linear map. Suppose that Y is of finite dimension. Then T
is bounded if and only if ker T := {x ∈ X : T x = 0} is closed.
Remark 3.9. The converse of PCorollary 3.8 does not hold when Y is of infinite dimension. For
∞
example, let X := {x ∈ ℓ : n=1 n |x(n)|2 < ∞} (note that X is a vector space Why?) and
2 2
Two normed spaces X and Y are said to be isomorphic (resp. isometric isomorphic) if there is
a bi-continuous linear isomorphism (resp. isometric) between X and Y . We write X = Y if X and
Y are isometric isomorphic.
Remark 3.10. Note that the inverse of a bounded linear isomorphism need not be bounded.
Example 3.11. Let X : {f ∈ C ∞ (−1, 1) : f (n) ∈ C b (−1, 1) for all n = 0, 1, 2...} and Y := {f ∈
X : f (0) = 0}. In addition, X and Y both are equipped with the sup-norm ∥ · ∥∞ . Define an
operator S : X → Y by Z x
Sf (x) := f (t)dt
0
for f ∈ X and x ∈ (−1, 1). Then S is a bounded linear isomorphism but its inverse S −1 is
unbounded. In fact, the inverse S −1 : Y → X is given by
S −1 g := g ′
for g ∈ Y .
A metric space is said to be separable if there is a countable dense subset, for example, the base
field K is separable. Moreover, it is easy to see that a normed space is separable if and only if it is
the closed linear span of a countable dense subset.
Definition 3.12. A sequence of element (en )∞ n=1 in a normed space X is called a Schauder basis
for X if for each element x ∈ X, there is a unique sequence of scalars (αn ) such that
∞
X
(3.2) x= αn e n .
n=1
Note: The expression in Eq. 3.2 depends on the order of en ’s.
Remark 3.13. Note that if X has a Schauder basis, then X must be separable. The following
natural question was first raised by Banach (1932).
The basis problem: Does every separable Banach space have a Schauder basis?
The answer is “No′′ !
This problem was completely solved by P. Enflo in 1973.
In the rest of this section, we are going to investigate some concrete examples of dual spaces.
Example 3.15. Let X = KN . Consider the usual Euclidean norm on X, i.e., ∥(x1 , ..., xN )∥ :=
|x1 |2 + · · · |xN |2 . Define θ : KN → (KN )∗ by θx(y) = x1 y1 + · · · + xN yN for x = (x1 , ..., xN )
p
and y = (y1 , ..., yN ) ∈ KN . Note that θx(y) = ⟨x, y⟩, the usual inner product on KN . Then by
the Cauchy-Schwarz inequality, it is easy to see that θ is an isometric isomorphism. Therefore, we
have KN = (KN )∗ .
10 CHI-WAI LEUNG
for x ∈ ℓ1 and η ∈ c0 .
Then T is isometric isomorphism and hence, c∗0 = ℓ1 .
Step 1 follows.
Step 2. T is an isometry.
Note that by Step 1, we have ∥T x∥ ≤ ∥x∥1 for all x ∈ ℓ1 . We need to show that ∥T x∥ ≥ ∥x∥1 for
all x ∈ ℓ1 . Fix x ∈ ℓ1 . Now for each k = 1, 2.., consider the polar form x(k) = |x(k)|eiθk . Note that
ηn := (e−iθ1 , ..., e−iθn , 0, 0, ....) ∈ c0 for all n = 1, 2.... Then we have
n
X n
X
|x(k)| = x(k)ηn (k) = T x(ηn ) = |T x(ηn )| ≤ ∥T x∥
k=1 k=1
Example 3.17. We have the other important examples of the dual spaces.
(i) (ℓ1 )∗ = ℓ∞ .
(ii) For 1 < p < ∞, (ℓp )∗ = ℓq , where p1 + 1q = 1.
(iii) For a locally compact Hausdorff space X, C0 (X)∗ = M (X), where M (X) denotes the space
of all regular Borel measures on X.
Parts (i) and (ii) can be obtained by the similar argument as in Example 3.16 (see also in [?,
Chapter 8]). Part (iii) is known as the Riesz representation Theorem which is referred to [?,
Section 21.5] for the details.
Example 3.18. Let C[a, b] be the space of all continuous R-valued functions defined on a closed
and bounded interval [a, b]. Moreover, the space C[a, b] is endowed with the sup-norm, i.e., ∥f ∥∞ :=
11
Let BV ([a, b]) denote the space of all bounded variations on [a, b] and let ∥ρ∥ := V (ρ) for ρ ∈
BV ([a, b]). Then BV ([a, b]) becomes a Banach space.
Besides, for f ∈ C[a, b], the Riemann-Stieltjes integral of f with respect to a bounded variation ρ
on [a, b] is defined by
Z b n
X
f (x)dρ(x) := lim f (ξk )(ρ(xk ) − ρ(xk−1 )),
a P
k=1
4. Hahn-Banach Theorem
A real valued function p : X → R defined on a vector space X is called a positively homogeneous
sub-additive if the following conditions hold:
(i) p(αx) = αp(x) for all x ∈ X and α ≥ 0.
(ii) p(x + y) ≤ p(x) + p(y) for all x, y ∈ X.
Lemma 4.1. Let X be a real vector space and Y be a subspace of X. Assume that there is an
element v ∈ X \ Y such that X = Y ⊕ Rv, i.e., the space X is the linear span of Y and v. Let p be
a positive homogeneous sub-additive function defined on X. Suppose that f is real linear functional
defined on Y satisfying f (y) ≤ p(y) for all y ∈ Y . Then there is a real linear extension F of f
defined on X so that
F (x) ≤ p(x) for all x ∈ X.
Proof. It is noted that if F is a linear extension of f on X and γ := F (v) which satisfies
F (y + tv) = f (y) + tγ ≤ p(y + tv) for all y ∈ Y and for all t ∈ R,
then it suffices to saying that the following inequalities hold:
(4.1) f (y1 ) + γ ≤ p(y1 + v) and f (y2 ) − γ ≤ p(y2 − v)
for all y1 , y2 ∈ Y . Thus, we need to determine γ := F (v) so that the following holds:
Remark 4.2. Before completing the proof of the Hahn-Banach Theorem, Let us first recall one
of super important results in mathematics, called Zorn’s Lemma, a very humble name. Every
mathematics student should know it.
Zorn’s Lemma: Let X be a non-empty set with a partially order “ ≤ ”. Assume that every totally
order subset C of X has an upper bound, i.e. there is an element z ∈ X such that c ≤ z for all c ∈ C.
Then X must contain a maximal element m, that is, if m ≤ x for some x ∈ X, then m = x.
Theorem 4.3. Hahn-Banach Theorem : Let X be a vector space ( not necessary to be a normed
space) over R and let Y be a subspace of X. Let p be a positive homogeneous sub-additive function
defined on X. Suppose that f is a real linear functional defined on Y satisfying f (y) ≤ p(y) for all
y ∈ Y . Then there is a real linear extension F of f defined on X so that
F (x) ≤ p(x) for all x ∈ X.
Proof. Let X be the collection of the pairs (Y1 , f1 ), where Y ⊆ Y1 is a subspace of X and f1 is a
linear extension of f defined on Y1 such that and f1 ≤ p on Y1 . Define a partial order ≤ on X by
(Y1 , f1 ) ≤ (Y2 , f2 ) if Y1 ⊆ Y2 and f2 |Y1 = f1 . Then by the Zorn’s lemma, there is a maximal element
(Ye , F ) in X. The maximality of (Ye , F ) and Lemma 4.1 give Ye = X. The proof is complete. □
Definition 4.4. Let D be a convex subset of a normed space X, i.e., tx + (1 − t)y ∈ D for all
x, y ∈ D and t ∈ (0, 1). Suppose that 0 is an interior point of D. Define
µD (x) := inf{t > 0 : x ∈ tD}
for x ∈ X. In addition, set µD (x) = ∞ if {t > 0 : x ∈ tD} = ∅.
The function µD is called the Minkowski functional with respect to D.
Lemma 4.5. Let D be a convex subset of a normed space X. Suppose that 0 is an interior point
of D. Then the Minkowski functional µ := µD : X → [0, ∞) is positively homogeneous and sub-
additive on D.
In addition, we have {x ∈ X : µ(x) < 1} ⊆ D ⊆ {x ∈ X : µ(x) ≤ 1}.
Proof. It is noted that since 0 ∈ int(D), the set {t > 0 : x ∈ tD} = ̸ ∅ for all x ∈ X. Thus, the
function µ : X → [0, ∞) is defined.
Clearly, if we fix t > 0 and x ∈ X, then we have µ(tx) ≤ s if and only if tµ(x) ≤ s. Hence, the
function µ is positively homogeneous.
Next, we show the subadditivity of µ. Let ε > 0. For x, y ∈ X, we choose s, t > 0 such that x ∈ sD
and y ∈ tD satisfying s < µ(x) + ε and t < µ(y) + ε. Then x = sd1 and y = td2 for some d1 , d2 ∈ D.
Since D is convex, we have
s t
x + y = sd1 + td2 = (s + t)( d1 + d2 ) ∈ (s + t)D.
s+t s+t
Thus, µ(x + y) ≤ s + t and so, µ(x + y) < µ(x) + µ(y) + 2ε. Therefore, µ is sub-additive. The last
assertion is clear by the definition of µ. □
13
Proposition 4.6. Let C be a closed convex subset of a real vector space X. Let d and A be the
positive constants. If x0 is an element in X with ∥x0 ∥ ≤ A such that 0 < d < dist(x0 , C), then
there is an element F1 ∈ BX ∗ such that
(4.3) F1 (y) + α < F1 (x0 ).
d d −1
for all y ∈ C, whenever 0 < α < − 2A
2 (1 ) .
Consequently, for any closed convex subset C1 of X and x′0 ∈
/ C1 , then there is an element g ∈ X ∗
with ∥g∥ ≤ 1 such that
(4.4) g(z) + β < g(x′0 ).
for all z ∈ C1 , whenever 0 < β < dist(x′0 , C1 ).
Proof. For showing the first assertion, we first note that if there is x0 ∈ X such that ∥x0 ∥ ≤ A and
d
d(x0 , C) > d. then 2A < 1. To see this, we have A ≥ ∥x0 ∥ ≥ d(x0 , C) > d2 because 0 ∈ C.
Now since 0 < d < dist(x0 , C), we have (x0 + B(0, d)) ∩ C = ∅. Thus, we have (x0 + B(0, 12 d)) ∩
(C + B(0, 21 d)) = ∅. Put D := C + B(0, 21 d). Notice that D is a convex subset of X and x0 ∈ / D.
Moreover, we have 0 ∈ int(D). Let µ := µD be the Minkowski functional corresponding to D.
Then µ is positive homogeneous and sub-additive on X by Lemma 4.5.
Put Y := Rx0 and define f : Y → R by f (αx0 ) := αµ(x0 ) for α ∈ R. Then f (y) ≤ µ(y) for all
y ∈ Y since µ ≥ 0 and positive homogenous. The Hahn-Banch Theorem 4.3 implies that there is a
linear extension F defined on X satisfying F (x) ≤ µ(x) for all x ∈ X. We want to show that the
linear functional F1 := d2 F ∈ BX ∗ is as required.
We first notice that F is bounded because we have |F (y)| ≤ µ(y) ≤ 1 for all y ∈ B(0, 12 d) ⊆ D and
so, ∥F1 ∥ = ∥ d2 F ∥ ≤ 1. Note that µ(x) ≤ 1 for all x ∈ C because C ⊆ D. Thus, sup F (C) ≤ 1.
On the other hand, since x0 ∈ / D, we have F (x0 ) = µ(x0 ) ≥ 1. Now if µ(x0 ) = 1, then there
is a decreasing sequence of positive numbers (λn ) with λn ↓ 1 and λ1n x0 ∈ D. This implies
that x0 ∈ D. It contradicts to the fact that (x0 + B(0, 21 d)) ∩ D is empty. Hence, we have
F (y) ≤ 1 < F (x0 ) = µ(x0 ) for all y ∈ D.
Next, we are going to show that the Inequality 4.3 holds. In fact, for λ > 0, we see that x0 ∈ λD
if and only if λ1 x0 ∈ D. Hence, In this case, we have 1 < µ(x0 ) ≤ λ. Also, we have λ1 x0 = y + z for
some y ∈ C and for some z ∈ B(0, 12 d). Then we have
1 1 1 1
d ≤ ∥x0 − y∥ = ∥x0 − x0 − z∥ = |1 − |∥x0 ∥ + ∥z∥ < |1 − |A + d.
λ λ λ 2
1 d d −1
This implies that 1 − λ > 2A because λ ≥ µ(x0 ) > 1. This gives 1 < (1 − 2A ) < λ whenever
d −1 d −1
λ > 0 with x0 ∈ λD and hence, µ(x0 ) ≥ (1 − 2A ) . Now if we put 0 < α1 := (1 − 2A ) − 1, then
we have
d −1
F (y) + α1 ≤ 1 + α1 < (1 − ) ≤ µ(x0 ) = F (x0 )
2A
for all y ∈ C. Therefore, if 0 < α < d2 α1 , then the element F1 := d2 F ∈ BX ∗ satisfies the inequality
4.3 as desired.
td0 td0 −1
For showing the last assertion, let d0 := dist(x′0 , C1 ) > 0. Clearly, we have lim (1− ) = d0 .
t→1− 2 2d0
Thus if 0 < β < d0 , there is 0 < t1 < 1 such that β < t12d0 (1 − t2d 1 d0 −1
0
) . Now if we put d := t1 d0 ,
then we have 0 < β < 2 (1 − 2d0 ) . From this, we choose ε > 0 such that β < d2 (1 − 2(d0d+ε) )−1 .
d d −1
Theorem 4.7. Let X be a normed space and let Y be a subspace of X. If f ∈ Y ∗ , then there exists
a linear extension F ∈ X ∗ of f such that ∥F ∥ = ∥f ∥.
Proof. W.L.O.G, we may assume that ∥f ∥ = 1. We first show the case when X is normed space
over R. It is noted that the norm function p(·) := ∥ · ∥ is positively homogeneous and sub-additive
on X. Since ∥f ∥ = 1, we have f (y) ≤ p(y) for all y ∈ Y . Then by the Hahn-Banach Theorem 4.3,
there is a linear extension F of f on X such that F (x) ≤ p(x) for all x ∈ X. This implies that
∥F ∥ = 1 as required.
Now for the complex case, let h = Ref and g = Imf . Then f = h + ig and f, g both are real
linear on Y with ∥h∥ ≤ 1. Note that since f (iy) = if (y) for all y ∈ Y , we have g(y) = −h(iy)
for all y ∈ Y . This gives f (·) = h(·) − ih(i·) on Y . Then by the real case above, there is a real
linear extension H on X such that ∥H∥ = ∥h∥. Now define F : X −→ C by F (·) := H(·) − iH(i·).
Then F ∈ X ∗ and F |Y = f . Thus it remains to show that ∥F ∥ = ∥f ∥ = 1. We need to show that
|F (z)| ≤ ∥z∥ for all z ∈ X. For z ∈ X, consider the polar form F (z) = reiθ . Then F (e−iθ z) = r ∈ R
and thus F (e−iθ z) = H(e−iθ z). This yields that
|F (z)| = r = |F (e−iθ z)| = |H(e−iθ z)| ≤ ∥H∥∥e−iθ z∥ ≤ ∥z∥.
The proof is complete. □
Proposition 4.8. Let X be a normed space and x0 ∈ X. Then there is f ∈ X ∗ with ∥f ∥ = 1 such
that f (x0 ) = ∥x0 ∥. Consequently, we have
∥x0 ∥ = sup{|g(x)| : g ∈ BX ∗ }.
In addition, if x, y ∈ X with x ̸= y, then there exists f ∈ X ∗ such that f (x) ̸= f (y).
Proof. Let Y = Kx0 . Define f0 : Y → K by f0 (αx0 ) := α∥x0 ∥ for α ∈ K. Then f0 ∈ Y ∗ with
∥f0 ∥ = ∥x0 ∥. The result follows immediately from the Hahn-Banach Theorem. □
Remark 4.9. Proposition 4.8 tells us that the dual space X ∗ of X must be non-zero. Indeed, the
dual space X ∗ is very “Large′′ so that it can separate any pair of distinct points in X.
Furthermore, for any normed space Y and any pair of points x1 , x2 ∈ X with x1 ̸= x2 , we can find
an element T ∈ B(X, Y ) such that T x1 ̸= T x2 . In fact, fix a non-zero element y ∈ Y . Then by
Proposition 4.8, there is f ∈ X ∗ such that f (x1 ) ̸= f (x2 ). Thus, if we define T x = f (x)y, then
T ∈ B(X, Y ).
Proposition 4.10. Using the notations as above, if M is closed subspace and v ∈ X \ M , then
there is f ∈ X ∗ such that f (M ) ≡ 0 and f (v) ̸= 0.
Proof. Since M is a closed subspace of X, we can consider the quotient space X/M . Let π : X →
X/M be the natural projection. Note that v̄ := π(v) ̸= 0 ∈ X/M because v̄ ∈ X \ M . Then by
Corollary 4.8, there is a non-zero element f¯ ∈ (X/M )∗ such that f¯(v̄) ̸= 0. Therefore, the linear
functional f := f¯ ◦ π ∈ X ∗ is as desired. □
Therefore, B(f, r) ∩ F is infinite for all r > 0. In this case, there is a subsequence (fnk ) such that
∥fnk − f ∥ → 0. This gives
1
∥fnk ∥ ≤ |fnk (xnk )| = |fnk (xnk ) − f (xnk )| ≤ ∥fnk − f ∥ → 0
2
because f (M ) ≡ 0. Thus∥fnk ∥ → 0 and hence f = 0. It leads to a contradiction again. Thus, we
can conclude that M = X as desired. □
Remark 4.12. The converse of Proposition 4.11 does not hold. For example, consider X = ℓ1 .
Then ℓ1 is separable but the dual space (ℓ1 )∗ = ℓ∞ is not.
Proposition 4.13. Let X and Y be normed spaces. For each element T ∈ B(X, Y ), define a linear
operator T ∗ : Y ∗ → X ∗ by
T ∗ y ∗ (x) := y ∗ (T x)
for y ∗ ∈ Y ∗ and x ∈ X. Then T ∗ ∈ B(Y ∗ , X ∗ ) and ∥T ∗ ∥ = ∥T ∥. In this case, T ∗ is called the
adjoint operator of T .
Proof. We first claim that ∥T ∗ ∥ ≤ ∥T ∥ and hence, ∥T ∗ ∥ is bounded.
In fact, for any y ∗ ∈ Y ∗ and x ∈ X, we have |T ∗ y ∗ (x)| = |y ∗ (T x)| ≤ ∥y ∗ ∥∥T ∥∥x∥. Hence,
∥T ∗ y ∗ ∥ ≤ ∥T ∥∥y ∗ ∥ for all y ∗ ∈ Y ∗ . Thus, ∥T ∗ ∥ ≤ ∥T ∥.
We need to show ∥T ∥ ≤ ∥T ∗ ∥. Let x ∈ BX . Then by Proposition 4.8, there is y ∗ ∈ SX ∗ such that
∥T x∥ = |y ∗ (T x)| = |T ∗ y ∗ (x)| ≤ ∥T ∗ y ∗ ∥ ≤ ∥T ∗ ∥. This implies that ∥T ∥ ≤ ∥T ∗ ∥. □
Example 4.14. Let X and Y be the finite dimensional normed spaces. Let (ei )ni=1 and (fj )m j=1 be
the bases for X and Y respectively. Let θX : X → X ∗ and θY : X → Y ∗ be the identifications as
in Example 3.15. Let e∗i := θX ei ∈ X ∗ and fj∗ := θY fj ∈ Y ∗ . Then e∗i (el ) = δil and fj∗ (fl ) = δjl ,
where, δil = 1 if i = l; otherwise is 0.
Now if T ∈ B(X, Y ) and (aij )m×n is the representative matrix of T corresponding to the bases
(ei )ni=1 and (fj )m ∗ ∗ ∗ ′
j=1 respectively, then akl = fk (T el ) = T fk (el ). Therefore, if (alk )n×m is the
representative matrix of T ∗ corresponding to the bases (fj∗ ) and (e∗i ), then akl = a′lk . Hence the
transpose (akl )t is the the representative matrix of T ∗ .
Proposition 4.15. Let Y be a closed subspace of a normed space X. Let i : Y → X be the natural
inclusion and π : X → X/Y the natural projection. Then
(i) the adjoint operator i∗∗ : Y ∗∗ → X ∗∗ is an isometry.
(ii) the adjoint operator π ∗ : (X/Y )∗ → X ∗ is an isometry.
Consequently, Y ∗∗ and (X/Y )∗ can be viewed as the closed subspaces of X ∗∗ and X ∗ respectively.
Proof. For Part (i), we first note that for any x∗ ∈ X ∗ , the image i∗ x∗ in Y ∗ is just the restriction
of x∗ on Y , denoted by x∗ |Y . Now let ϕ ∈ Y ∗∗ . Then for any x∗ ∈ X ∗ , we have
|i∗∗ ϕ(x∗ )| = |ϕ(i∗ x∗ )| = |ϕ(x∗ |Y )| ≤ ∥ϕ∥∥x∗ |Y ∥Y ∗ ≤ ∥ϕ∥∥x∗ ∥X ∗ .
Thus, ∥i∗∗ ϕ∥ ≤ ∥ϕ∥. WE need to show the inverse inequality. Now for each y ∗ ∈ Y ∗ , the Hahn-
Banach Theorem gives an element x∗ ∈ X ∗ such that ∥x∗ ∥X ∗ = ∥y ∗ ∥Y ∗ and x∗ |Y = y ∗ and hence,
i∗ x∗ = y ∗ . Then we have
|ϕ(y ∗ )| = |ϕ(x∗ |Y )| = |ϕ(i∗ x∗ )| = |(i∗∗ ◦ ϕ)(x∗ )| ≤ ∥i∗∗ ϕ∥∥x∗ ∥X ∗ = ∥i∗∗ ϕ∥∥y ∗ ∥Y ∗
for all y ∗ ∈ Y ∗ . Therefore, we have ∥i∗∗ ϕ∥ = ∥ϕ∥.
For Part (ii), let ψ ∈ (X/Y )∗ . Note that since ∥π ∗ ∥ = ∥π∥ ≤ 1, we have ∥π ∗ ψ∥ ≤ ∥ψ∥. On the
other hand, for each x̄ := π(x) ∈ X/Y with ∥x̄∥ < 1, we can choose an element m ∈ Y such that
∥x + m∥ < 1. Therefore, we have
|ψ(x̄)| = |ψ ◦ π(x)| = |ψ ◦ π(x + m)∥ ≤ ∥ψ ◦ π∥ = ∥π ∗ (ψ)∥.
16 CHI-WAI LEUNG
Remark 4.16. By using Proposition 4.15, we can give an alternative proof of the Riesz’s Lemma
2.5.
Using the notations as in Proposition 4.15, if Y ⊊ X, then we have ∥π∥ = ∥π ∗ ∥ = 1 because π ∗ is an
isometry by Proposition 4.15(ii). Thus we have ∥π∥ = sup{∥π(x)∥ : x ∈ X, ∥x∥ = 1} = 1. Hence,
for any 0 < θ < 1, we can find element z ∈ X with ∥z∥ = 1 such that θ < ∥π(z)∥ = inf{∥z + y∥ :
y ∈ Y }. The Riesz’s Lemma follows.
5. Reflexive Spaces
Proposition 5.1. For a normed space X, let Q : X −→ X ∗∗ be the canonical map, that is,
Qx(x∗ ) := x∗ (x) for x∗ ∈ X ∗ and x ∈ X. Then Q is an isometry.
Proof. Note that for x ∈ X and x∗ ∈ BX ∗ , we have |Q(x)(x∗ )| = |x∗ (x)| ≤ ∥x∥. Then ∥Q(x)∥ ≤
∥x∥.
We need to show that ∥x∥ ≤ ∥Q(x)∥ for all x ∈ X. In fact, for x ∈ X, there is x∗ ∈ X ∗ with
∥x∗ ∥ = 1 such that ∥x∥ = |x∗ (x)| = |Q(x)(x∗ )| by Proposition 4.8. Thus we have ∥x∥ ≤ ∥Q(x)∥.
The proof is complete. □
Remark 5.2. Let T : X → Y be a bounded linear operator and T ∗∗ : X ∗∗ → Y ∗∗ the second dual
operator induced by the adjoint operator of T . Using notations as in Proposition 5.1 above, the
following diagram commutes.
T
X −−−−→ Y
QX y
Q
y Y
T ∗∗
X ∗∗ −−−−→ Y ∗∗
Example 5.6. By using Proposition 5.5, we immediately see that the space ℓ∞ is not reflexive
because it contains a non-reflexive closed subspace c0 .
Proposition 5.7. Let X be a Banach space. Then we have the following assertions.
(i) X is reflexive if and only if the dual space X ∗ is reflexive.
(ii) If X is reflexive, then so is every quotient of X.
Proof. For Part (i), suppose that X is reflexive first. Let ze ∈ X ∗∗∗ . Then the restriction z := ze|X ∈
X ∗ . Then one can directly check that Qz = z on X ∗∗ since X ∗∗ = X.
For the converse, assume that X ∗ is reflexive but X is not. Therefore, X is a proper closed subspace
of X ∗∗ . Then by using the Hahn-Banach Theorem, we can find a non-zero element ϕ ∈ X ∗∗∗ such
that ϕ(X) ≡ 0. However, since X ∗∗∗ is reflexive, we have ϕ ∈ X ∗ and hence, ϕ = 0 which leads to
a contradiction.
For Part (ii), we assume that X is reflexive. Let M be a closed subspace of X and π : X → X/M
the natural projection. Note that the adjoint operator π ∗ : (X/M )∗ → X ∗ is an isometry (Check
!). Thus, (X/M )∗ can be viewed as a closed subspace of X ∗ . By Part (i) and Proposition 5.5, we
see that (X/M )∗ is reflexive. Then X/M is reflexive by using Part (i) again.
The proof is complete. □
Lemma 5.8. Let M be a closed subspace of a normed space X. Let r : X ∗ → M ∗ be the restriction
map, that is x∗ ∈ X ∗ 7→ x∗ |M ∈ M ∗ . Put M ⊥ := ker r := {x∗ ∈ X ∗ : x∗ (M ) ≡ 0}. Then the
canonical linear isomorphism re : X ∗ /M ⊥ → M ∗ induced by r is an isometric isomorphism.
Proof. We first note that r is surjective by using the Hahn-Banach Theorem. We need to show
that re is an isometry. Note that re(x∗ + M ⊥ ) = x∗ |M for all x∗ ∈ X ∗ . Now for any x∗ ∈ X ∗ , we
have ∥x∗ + y ∗ ∥X ∗ ≥ ∥x∗ + y ∗ ∥M ∗ = ∥x∗ |M ∥M ∗ for all y ∗ ∈ M ⊥ . Thus, we have ∥e
r(x∗ + M ⊥ )∥ =
∗ ∗ ⊥
∥x |M ∥M ∗ ≤ ∥x + M ∥. We need to show the reverse inequality.
Now for any x∗ ∈ X ∗ , then by the Hahn-Banach Theorem again, there is z ∗ ∈ X ∗ such that
z ∗ |M = x∗ |M and ∥z ∗ ∥ = ∥x∗ |M ∥M ∗ . Then x∗ − z ∗ ∈ M ⊥ and hence, we have x∗ + M ⊥ = z ∗ + M ⊥ .
This implies that
∥x∗ + M ⊥ ∥ = ∥z ∗ + M ⊥ ∥ ≤ ∥z ∗ ∥ = ∥x∗ |M ∥M ∗ = ∥e
r(x∗ + M ⊥ )∥.
The proof is complete. □
Remark 5.10. In view of the definition of a reflexive space, it is naturally raised the question
that whether a Banach space X is reflexive whenever it is isometrically isomorphic to its second
dual. The answer is negative. A counter example was given by R.C. James in 1951 (see [8]).
Proposition 6.2. A weak limit of a sequence is unique if it exists. In this case, if (xn ) weakly
w
converges to x, denoted by x = w-lim xn or xn −
→ x.
n
w
Remark 6.3. Clearly, if a sequence (xn ) converges to x ∈ X in norm, then xn −
→ x. However,
the weakly convergence of a sequence does not imply the norm convergence.
For example, consider X = c0 and (en ). Then f (en ) → 0 for all f ∈ c∗0 = ℓ1 but (en ) is not
convergent in c0 .
Proposition 6.4. Suppose that X is finite dimensional. A sequence (xn ) in X is norm convergent
if and only if it is weakly convergent.
Proof. Suppose that (xn ) weakly converges to x. Let B := {e1 , .., eN } be a basis for X and let fk be
the k-th coordinate functional corresponding to the basis B, i.e., v = N
P
k=1 fk (v)ek for all v ∈ X.
Since dim X < ∞, we have fk in X ∗ for all k = 1, ..., N . Therefore, we have limn fk (xn ) = fk (x)
for all k = 1, ..., N . Thus, we have ∥xn − x∥ → 0. □
Definition 6.5. Let X be a normed space. A sequence (fn ) in X ∗ is said to be weak∗ convergent
if there is f ∈ X ∗ such that limn fn (x) = f (x) for all x ∈ X, that is fn point-wise converges to f .
w∗
In this case, f is called the weak∗ limit of (fn ). Write f = w∗ -limn fn or fn −−→ f .
Remark 6.6. In the dual space X ∗ of a normed space X, we always have the following implications:
“Norm Convergent” =⇒ “Weakly Convergent” =⇒ “Weak∗ Convergent”.
However, the converse of each implication does not hold.
19
Example 6.7. Remark 6.3 has shown that the w-convergence does not imply ∥ · ∥-convergence.
We now claim that the w∗ -convergence also Does Not imply the w-convergence.
Consider X = c0 . Then c∗0 = ℓ1 and c∗∗ 1 ∗ ∞ ∗ 1 ∗
0 = (ℓ ) = ℓ . Let en = (0, ...0, 1, 0...) ∈ ℓ = c0 , where
w∗
the n-th coordinate is 1. Then e∗n −−→ 0 but e∗n ↛ 0 weakly because e∗∗ (e∗n ) ≡ 1 for all n, where
e∗∗ := (1, 1, ...) ∈ ℓ∞ = c∗∗ ∗
0 . Hence the w -convergence does not imply the w-convergence.
w
Proposition 6.8. Let (fn ) be a sequence in X ∗ . Suppose that X is reflexive. Then fn −
→ f if and
w∗
only if fn −−→ f .
In particular, if dim X < ∞, then the followings are equivalent:
∥·∥
(i) : fn −−→ f ;
w
(ii) : fn −→ f;
w∗
(iii) : fn −−→ f .
Theorem 6.9. (Banach) : Let X be a separable normed space. If (fn ) is a bounded sequence in
X ∗ , then it has a w∗ -convergent subsequence.
Proof. Let D := {x1 , x2 , ...} be a countable dense subset of X. Note that since (fn )∞ n=1 is bounded,
(fn (x1 )) is a bounded sequence in K. Then (fn (x1 )) has a convergent subsequence, say (f1,k (x1 ))∞k=1
in K. Let c1 := limk f1,k (x1 ). Now consider the bounded sequence (f1,k (x2 )). Then there is
convergent subsequence, say (f2,k (x2 )), of (f1,k (x2 )). Put c2 := limk f2,k (x2 ). Note that we still
have c1 = limk f2,k (x1 ). To repeat the same step, if we define (m, k) ≤ (m′ , k ′ ) if m < m′ ; or
m = m′ with k ≤ k ′ , we can find a sequence (fm,k )m,k in X ∗ such that
(i) : (fm+1,k )∞ ∞
k=1 is a subsequence of (fm,k )k=1 for m = 0, 1, .., where f0,k := fk .
(ii) : ci = limk fm,k (xi ) exists for all 1 ≤ i ≤ m.
Now put hk := fk,k . Then (hk ) is a subsequence of (fn ). Note that for each i, we have limk hk (xi ) =
limk fi,k (xi ) = ci by the construction (ii) above. Since (∥hk ∥) is bounded and D is dense in X, we
have h(x) := limk hk (x) exists for all x ∈ X and h ∈ X ∗ . That is h = w∗ -limk hk . The proof is
complete. □
Remark 6.10. Theorem 6.9 does not hold if the separability of X is removed.
For example, consider X = ℓ∞ and δn the n-th coordinate functional on ℓ∞ . Then δn ∈ (ℓ∞ )∗
with ∥δn ∥(ℓ∞ )∗ = 1 for all n. Suppose that (δn ) has a w∗ -convergent subsequence (δnk )∞
k=1 . Define
x ∈ ℓ∞ by
0
if m ̸= nk ;
x(m) = 1 if m = n2k ;
−1 if m = n2k+1 .
Hence we have |δni (x) − δni+1 (x)| = 2 for all i = 1, 2, ... It leads to a contradiction. Thus(δn ) has
no w∗ -convergent subsequence.
Corollary 6.12. Suppose that X is a separable. If X is reflexive space, then the closed unit ball
BX of X is sequentially weakly compact, i.e. it is equivalent to saying that any bounded sequence
in X has a weakly convergent subsequence.
Proof. Let Q : X → X ∗∗ be the canonical map as before. Let (xn ) be a bounded sequence in X.
Hence, (Qxn ) is a bounded sequence in X ∗∗ . We first note that since X is reflexive and separable,
X ∗ is also separable by Proposition 4.11. We can apply Theorem 6.9, (Qxn ) has a w∗ -convergent
subsequence (QxnK ) in X ∗∗ = Q(X) and hence, (xnk ) is weakly convergent in X. □
Remark 6.13. In fact, the converse of Corollary 6.12 also holds (see Appendix 7 below). The
assumption of separability of X can be removed. We have the following stronger result which was
shown by R. C. James (see [10, §1.13]).
Theorem 6.14. Let X be a Banach space. Then the following are equivalent.
(i) X is reflexive.
(ii) Every bounded sequence in X has a weakly convergent subsequence.
(iii) The closed unit ball BX of X is weakly compact, that is, BX is compact in the weak topology.
Before showing the main theorem, we need some basic knowledge of the set theory which can be
found in the Halmos’s classic book [7].
Recall that a partially order set S is called a well ordered set if for any non-empty subset A of S
contains the least element, that is, there is an element x0 ∈ A such that x0 ≤ x for all x ∈ A. In
particular, S is automatically a totally order set. The well-ordering theorem tells us that every
set can be equipped with a well-ordering.
Two well ordered sets A and B are said to have the same ordinal number or ordinal for simply
if there is an order preserving bijection from A onto B. In particular, each ordinal can be viewed
as a well ordered set. More precisely, an ordinal number α is a well ordered set such that for any
element η ∈ α, we have η = {ξ ∈ α : ξ < η}, thus, we have η ⊆ α whenever η ∈ α. This definition
was due to von Neumann.
Put ω the least infinite countable ordinal, that is, ω := {0, 1, 2...} and is endowed with the usual
order.
On the other hand, it is naturally led to define an order on the class of ordinals as the following.
(Warning: We DO NOT HAVE a statement about ” the set of All ordinals numbers” !!!! )
Definition 7.1. Let α and β be two ordinals. We say that
(i) α = β if there is an order preserving bijection from α onto β.
(ii) α ≤ β if there is an order preserving injection from α to β.
(iii) α < β if α ≤ β but α ̸= β.
From the von Neumann’s definition, we have (see [7, Section 20])
21
Lemma 7.2. Every nonempty set of ordinal numbers is a well ordered set.
Next we need the following definition for comparing the size of two given sets.
Definition 7.3. Two sets A and B are said to have the same cardinality, write A ≈ B, if there is
a bijection from A onto B.
The cardinal number of A, write |A|, is defined by an ordinal number given by
|A| := min{α : α is an ordinal number such that α ≈ A}
(Notice that the well ordering theorem and Lemma 7.2 assure the existence of |A|. )
exists.
(iv) Let θ be a limit ordinal. A subset M of θ is said to be cofinal if for every ordinal µ < θ,
there is ν ∈ M such that µ < ν < θ. In this case, if lim xξ exists, then so does lim xν
ξ→θ ν∈M :ν→θ
and they are the same.
Lemma 7.6. Let X be a Banach space. If (fξ )ξ<θ is a norm bounded θ-sequence in X ∗ , then there
is an element f ∈ X ∗ such that ∥f ∥ ≤ supξ<θ ∥fξ ∥ and
for all x ∈ X.
In this case, f is called a transfinite limit of (fξ ) ( note that f may not be unique).
22 CHI-WAI LEUNG
Proof. Let M := supξ<θ ∥fξ ∥. We first notice that since (fξ )ξ<θ is bounded, lim fξ (x) exists for all
ξ→θ
x ∈ X. Hence, one can define a function p : X → R by
p(x) := lim fξ (x)
ξ→θ
for x ∈ X. Clearly, p is a positively homogenous and sub-additive function. We may assume that
p(x0 ) > 0 for some x0 ∈ X. To see this, if p(x0 ) < 0, then p(−x0 ) = limfξ (−x0 ) ≥ limfξ (−x0 ) =
−limfξ (x0 ) > 0 as desired. Now if we define a linear map f0 on Rx0 by f0 (tx0 ) := tp(x0 ), then
f0 (tx0 ) ≤ p(tx0 ) for all t ∈ R. Then by the Hahn-Banach Theorem 4.3, there is a linear extension
f of f0 defined on X such that f (x) ≤ p(x) for all x ∈ X. Notice that since |fξ (x)| ≤ M ∥x∥ for
all ξ < θ and for all x ∈ X, we have p(x) ≤ M ∥x∥. Thus, we have ∥f ∥ ≤ M as desired. The last
assertion is obtained by putting −x ∈ X into Eq 7.1. The proof is complete. □
Definition 7.7. A normed subspace Γ of X ∗ is said to be transfinitely closed if for every norm
bounded transfinite θ-sequence (fξ )ξ<θ in Γ for some limit ordinal θ, one can find an element f ∈ Γ
satisfying the Eq 7.1 above, that is,
f (x) ≤ lim fξ (x)
ξ→θ
for all x ∈ X.
Clearly, every transfinitely closed subspace is norm closed by considering θ = ω in Eq 7.2 above.
Recall the notation that for every element x in a Banach space X, x b denotes the element in X ∗∗
given by x(f ) := f (x) for all f ∈ X ∗ . Put A a : a ∈ A} ⊆ X ∗∗ for a subset A of X.
b := {b
Lemma 7.10. Let X be a Banach space. Assume that every normed bounded sequence in X has
a weakly convergent subsequence. Let D be a countably infinite subset of X. Then every bounded
transfinite sequence in D b has a transfinite limit in X x : x ∈ X} ⊆ X ∗∗ , that is, for every
b := {b
transfinite sequence (xξ )ξ<θ in D, there is an element z ∈ X such that
(7.6) f (z) ≤ lim f (xξ )
ξ→θ
for all f ∈ X ∗ .
Proof. Let θ be a limit ordinal and (xξ )ξ<θ be a bounded transfinite sequence in D. By the
assumption of D, we can write D = {xi : i = 0, 1, 2...}.
Case 1: there is an infinite sequence (ξn )∞
n=0 in [0, θ) such that for every ordinal η < θ, there is
N ∈ N such that η < ξn < θ for all n > N , write lim ξn = θ. In this case, put xξi := xi for
n→∞
i = 0, 1, 2.... Then by assumption, there is a weakly convergent sequence (xξik ) with the weak limit,
say z ∈ X. This implies that
zb(f ) = f (z) = lim f (xξik ) ≤ lim f (xξ ) = lim xbξ (f )
k→∞ ξ→θ ξ→θ
24 CHI-WAI LEUNG
for all f ∈ X ∗ .
The proof is complete if Eq 7.6 also holds for the following case.
Case 2: there is no sequence (ξn ) in [0, θ) such that lim ξn = θ.
n→∞
In this case, for each xi ∈ D , let Zi := {ξ < θ : xξ = xi }. We will see that Zi is confinal subset
of [0, ∞) for some i ∈ ω, that is, for any ordinal η < θ, there is ν ∈ Zi such that η < ν < θ.
To see this, if we assume that every Zi , i ∈ ω, is not a cofinal subset of [0, θ), then for each iω,
there is an ordinal µi with µi < θ such that ξ ≤ µi for all ξ ∈ Zi . Now put λ0 := µ0 and
λn := max{µ : i = 0, 1, .., n − 1}. This gives an increasing sequence (λn )n∈ω in [0, θ). Then by the
assumption of thisS case, there is λ ∈ [0, θ) such that λn ≤ λ for all n ∈ ω and hence, ξ ≤ λ for all
ξ ∈ [0, θ) because i∈ω Zi = [0, θ). This implies that there is no ordinal ξ such thhat λ < ξ < θ
that will lead to a contradiction because θ is a limit ordinal.
Now let Zi0 be a confinal subset of [0, θ) and z := xi0 . Then by the definition of Zi0 , xξ = z for all
ξ ∈ Zi0 . Thus, we have
zb(f ) = f (z) = lim f (xξ ) ≤ lim f (xξ ) = lim xbξ (f ).
ξ∈Zi0 ;ξ→θ ξ→θ ξ→θ
We are now in a position to reach the following main result in this section.
Theorem 7.11. Let X be a separable Banach space. Then X is reflexive if and only if every
bounded sequence in X has a weakly convergent subsequence.
Proof. The necessary condition has been shown in Corollary 6.12.
We are going to show the converse statement. Let D be a countable subset of X.
Claim: The space X b := {b x : x ∈ X} is transfinitely closed in X ∗∗ , that is, for every bounded
b there is an element z ∈ X such that
transfinite sequence (xbξ )ξ<θ in X,
f (z) ≤ lim f (xξ )
ξ→θ
for all f ∈ X ∗.
Now for each n = 1, 2... and each xξ , ξ < θ, we choose an element xnξ ∈ D such that
1
∥xξ − xnξ ∥ <
.
n
From this, for each n = 1, 2, ..., we obtain a bounded a θ-transfinite sequence (xnξ )ξ<θ in D. For
each n = 1, 2.., Lemma 7.10 gives an element zn ∈ X such that
f (zn ) ≤ lim f (xnξ )
ξ→θ
for all f ∈ X ∗. In addition from this we have ∥zn ∥ ≤ 1 + supξ<θ ∥xξ ∥ for all n = 1, 2, .... Then the
necessary condition implies that (zn ) has a weak convergent subsequence (znj ). Let z be the weak
limit of (znj ). Then we have
f (z) = lim f (znj )
j→∞
n
≤ lim lim f (xξ j )
j→∞ ξ→θ
∥f ∥
≤ lim ( lim f (xξ ) + )
j→∞ ξ→θ nj
= lim f (xξ )
ξ→θ
for all f ∈ X ∗.
The Claim follows.
b ⊊ X ∗∗ , then there is an element ϕ ∈ X ∗∗ \ X.
Finally, if X b Note that X
b is a closed subspace of
X . Using Lemma 7.9, the transfinite closeness of X implies that there is an element f0 ∈ X ∗
∗∗ b
25
b(f0 ) = 0 for all x ∈ X and so, f0 = 0 but ϕ(f0 ) = 1 that is ridiculous. Thus,
such that f0 (x) = x
Xb = X ∗∗ . The proof is complete. □
8. Appendix: w∗ -compactness
Throughout this section X always denotes a normed space. I suppose that the students have
learned a standard course of topology before.
Now for each ε > 0 and for finitely many elements x1 , ..., xm in X, let
W (x1 , .., xm ; ε) := {f ∈ X ∗ : |f (xi )| < ε; ∀i = 1, .., m}.
It is noted that 0 ∈ W (x1 , .., xm ; ε) for any ε > 0 and for all finitely many elements x1 , ..., xm in X.
Definition 8.1. The weak ∗ -topology on the dual space X ∗ is the topology generated by the collection
{h + W (x1 , .., xm ; ε) : h ∈ X ∗ ; for ε > 0 and for finitely many x1 , .., xm ∈ X}.
The following result is an important feature for a weak∗ -continuous linear functional on X ∗ .
2
(8.1) |φ(f )| ≤ L(f )
δ
whenever f ∈ X ∗ with L(f ) > 0.
Claim 1: Eq 8.1 holds for all f ∈ X ∗ . To see this, it suffices to show that if L(f ) = 0, then
φ(f ) = 0. In fact, if L(f ) = 0, then L(tf ) = tL(f ) = 0 for all t > 0. In particular, we have
|(tf )(xk )| = 0 < δ for all t > 0 and for all k = 1, .., N . This implies that t|φ(f )| = |φ(tf )| < 1 for
all t > 0. Thus, we have φ(f ) = 0 as desired.
So, from Eq 8.1, we have
N
\
(8.2) ck ⊆ ker φ.
ker x
k=1
Claim 2: φ ∈ span{c xk : k = 1, ..., N } in X ∗∗ . To see this, we first notice that if xbj ∈ span{c xk :
T TN T
k ̸= j} for some j, then k̸=j ker x ck ⊆ ker xbj and so, we have k=1 ker x ck = k̸=j ker x
ck . Therefore,
we may assume that {c x1 , ..., xcN } is a linearly independent set by considering a maximal linearly
independent subset of {c x1 , ..., xcN }. In this case, {x1 , ..., xN } is a linearly independent subset of X.
Then the Hahn-Banach Theorem implies that there are x∗1 , ..., x∗N in X ∗ such that x∗i (xj ) = 1 if
i = j; otherwise is 0 for i, j = 1, ..., N . From this for any f ∈ X ∗ , we have xbj (f − N ∗
P
PN k=1 f (xk )xk ) = 0
∗ ∗
for all j = 1, ..., N . Eq 8.2 implies that φ(f − k=1 f (xk )xk ) = 0 for all f ∈ X . This gives
N
X N
X
φ(f ) = f (xk )φ(x∗k ) = φ(x∗k )c
xk (f )
k=1 k=1
26 CHI-WAI LEUNG
Before showing the main result in this section, let us recall that product topologies.
Let (Zi )i∈I be a collection of topological spaces. Let Z be the usual Cartesian product, that is
Y [
Z := Zi : {z : I → Zi : z(i) ∈ Zi ; ∀i ∈ I}.
i∈I i∈I
Let pi : Z → Zi be the natural projection for i ∈ I. The product topology on Z is the weakest
topology such that each projection pi is continuous. More precisely, the following collection forms
an open basis for the product topology:
\
{ p−1 i (Wi ) : J is a finite subset of I and Wi is an open subset of Zi }.
i∈J
We have the following famous result in topology.
Theorem 8.4. Tychonoff ’s Theorem: The Cartesian product of compact spaces is compact
under the product topology.
Theorem 8.5. The closed unit ball BX ∗ of the dual space X is compact with respect to the weak ∗ -
topology.
Proof. For each x ∈ X, put Zx := [−∥x∥, ∥x∥] ⊆ R. Each Zx is endowed with the usual subspace
topology of R. Then Zx is a compact set for all x ∈ X. Let
Y
Z := Zx .
x∈X
Then the set Z is a compact Hausdorff space under the product topology. Define a mapping by
T : f ∈ BX ∗ 7→ T f ∈ Z; T f (x) := f (x) ∈ Zx for x ∈ X.
Then by the definitions of weak ∗ -topology and the product topology, it is clear that T is a home-
omorphism from BX ∗ onto its image T (BX ∗ ). Recall a fact that any closed subset of a compact
Hausdorff space is compact. Since Z is compact Haudorsff, it suffices to show that T (BX ∗ ) is a
closed subset of Z.
Let z ∈ T (BX ∗ ). We are going to show that there is an element f ∈ BX ∗ such that f (x) = z(x)
27
for all x ∈ X.
Define a function f : X → K by
f (x) := z(x)
for x ∈ X.
Claim : f (x+y) = f (x)+f (y) for all x, y ∈ X. In fact if we fix x, y ∈ X and for any ε > 0, then by
the definition of product topology, there is an element g ∈ BX ∗ such that |g(x + y) − z(x + y)| < ε;
|g(x) − z(x)| < ε; and |g(y) − z(y)| < ε. Since g is linear, we have g(x + y) − g(x) − g(y) = 0. This
implies that
|z(x + y) − z(x) − z(y)| = |z(x + y) − g(x + y) − (z(x) − g(x)) − (z(y) − g(y))| < 3ε
for all ε > 0. Thus we have z(x + y) = z(x) = z(y). The Claim follows.
Similarly, we have z(αx) = αz(x) for all α ∈ K and for all x ∈ X.
Therefore, the functional f (x) := z(x) is linear on X. It remains to show f is bounded with
∥f ∥ ≤ 1. In fact, for any x ∈ X and any ε > 0, then there is an element g ∈ BX ∗ such that
g(x) − z(x)| < ε. Therefore, we have |f (x)| = |z(x)| ≤ |g(x)| + ε ≤ ∥x∥ + ε. Therefore, f is bounded
and ∥f ∥ ≤ 1 as desired. The proof is complete. □
Lemma 9.2. Let X and Y be normed spaces and T : X → Y a linear map. Then T is open if and
only if 0 is an interior point of T (U ) where U is the open unit ball of X.
Proof. The necessary condition is obvious.
For the converse, let W be a non-empty subset of X and a ∈ W . Put b = T a. Since W is open,
we choose r > 0 such that BX (a, r) ⊆ W . Note that U = 1r (BX (a, r) − a) ⊆ 1r (W − a). Thus, we
have T (U ) ⊆ 1r (T (W ) − b). Then by the assumption, there is δ > 0 such that BY (0, δ) ⊆ T (U ) ⊆
1
r (T (W ) − b). This implies that b + rBY (0, δ) ⊆ T (W ) and so, T (a) = b is an interior point of
T (W ). □
Corollary 9.3. Let M be a closed subspace of a normed space X. Then the natural projection
π : X → X/M is an open map.
Proof. Put U and V the open unit balls of X and X/M respectively. Using Lemma 9.2, the result
is obtained by showing that V ⊆ π(U ). Note that if x̄ = π(x) ∈ V , then by the definition a quotient
norm, we can find an element m ∈ M such that ∥x + m∥ < 1. Hence we have x + m ∈ U and
x̄ = π(x + m) ∈ π(U ). □
Before showing the main result, we have to make use one of important properties of a metric
space which is known as the Baire Category Theorem. Recall that a subset A of a metric space E
is called a nowhere dense set if the closure A of A has no interior point.
Proposition 9.4. Let E be a complete metric space with a metric d. If E is a union of a sequence
of subsets (An ) of E, then int(AN ) ̸= ∅ for some AN . Hence, every complete metric space is not a
countable union of nowhere dense sets.
28 CHI-WAI LEUNG
Lemma 9.5. Let T : X −→ Y be a bounded linear surjection from a Banach space X onto a
Banach space Y . Then 0 is an interior point of T (U ), where U is the open unit ball of X, i.e.,
U := {x ∈ X : ∥x∥ < 1}.
Proof. Set U (r) := {x ∈ X : ∥x∥ < r} for r > 0 and so, U = U (1).
Claim 1 : 0 is an interior point of T (U (1)).
S∞
Note that since T is surjective, Y = n=1 T (U (n)). Then by the Baire Category Theorem, there
exists N such that int T (U (N )) ̸= ∅. Let y ′ be an interior point of T (U (N )). Then there is
η > 0 such that BY (y ′ , η) ⊆ T (U (N )). Since BY (y ′ , η) ∩ T (U (N )) ̸= ∅, we may assume that
y ′ ∈ T (U (N )). Let x′ ∈ U (N ) such that T (x′ ) = y ′ . Then we have
0 ∈ BY (y ′ , η) − y ′ ⊆ T (U (N )) − T (x′ ) ⊆ T (U (2N )) = 2N T (U (1)).
Thus, we have 0 ∈ 2N 1
(BY (y ′ , η) − y ′ ) ⊆ T (U (1)). Hence 0 is an interior point of T (U (1)). The
Claim 1 follows.
Therefore there is r > 0 such that BY (0, r) ⊆ T (U (1)). This implies that we have
(9.1) BY (0, r/2k ) ⊆ T (U (1/2k ))
for all k = 0, 1, 2....
Claim 2 : D := BY (0, r) ⊆ T (U (3)).
Let y ∈ D. By Eq 9.1, there is x1 ∈ U (1) such that ∥y − T (x1 )∥ < r/2. Then by using Eq 9.1
again, there is x2 ∈ U (1/2) such that ∥y − T (x1 ) − T (x2 )∥ < r/22 . To repeat the same steps, there
exists is a sequence (xk ) such that xk ∈ U (1/2k−1 ) and
∥y − T (x1 ) − T (x2 ) − ... − T (xk )∥ < r/2k
for all k. On the other hand, since ∞
P P∞ k−1 and X is Banach, x :=
P∞
k=1 ∥xk ∥ ≤ k=1 1/2 k=1 xk
exists in X and ∥x∥ ≤ 2. This implies that y = T (x) and ∥x∥ < 3.
Thus we the result follows. □
.
Theorem 9.6. Open Mapping Theorem : Using the notations as in Lemma 9.5, then T is an
open mapping.
Proof. The proof is complete by using Lemmas 9.2 and 9.5. □
Proposition 9.7. Let T be a bounded linear isomorphism between Banach spaces X and Y . Then
T −1 is bounded.
29
Consequently, if ∥ · ∥ and ∥ · ∥′ both are complete norms on X such that ∥ · ∥ ≤ c∥ · ∥′ for some c > 0,
then these two norms ∥ · ∥ and ∥ · ∥′ are equivalent.
Proof. The first assertion follows immediately from the Open Mapping Theorem.
Therefore, the last assertion can be obtained by considering the identity map I : (X, ∥·∥) → (X, ∥·∥′ )
which is bounded by the assumption. □
Corollary 9.8. Let X and Y be Banach spaces and T : X → Y a bounded linear operator. Then
the followings are equivalent.
(i) The image of T is closed in Y .
(ii) There is c > 0 such that
d(x, ker T ) ≤ c∥T x∥
for all x ∈ X.
(iii) If (xn ) is a sequence in X such that ∥xn + ker T ∥ = 1 for all n, then ∥T xn ∥ ↛ 0.
Proof. Let Z be the image of T . Then the canonical map Te : X/kerT → Z induced by T is a
bounded linear isomorphism. Note that Te(x̄) = T x for all x ∈ X, where x̄ := x + ker T ∈ X/ ker T .
For (i) ⇒ (ii): suppose that Z is closed. Then Z becomes a Banach space. Then the Open
Mapping Theorem implies that the inverse of Te is also bounded. Thus, there is c > 0 such that
d(x, ker T ) = ∥x̄∥X/ ker T ≤ c∥Te(x̄)∥ = c∥T (x)∥ for all x ∈ X. The part (ii) follows.
For (ii) ⇒ (i), let (xn ) be a sequence in X such that lim T xn = y ∈ Y exists and so, (T xn ) is
a Cauchy sequence in Y . Then by the assumption, (x̄n ) is a Cauchy sequence in X/ ker T . Since
X/ ker T is complete, we can find an element x ∈ X such that lim x̄n = x̄ in X/ ker T . This gives
y = lim T (xn ) = lim Te(x̄n ) = Te(x̄) = T (x). Therefore, y ∈ Z.
(ii) ⇔ (iii) is obvious. The proof is complete. □
Proposition 9.9. Let X and Y be Banach spaces. Let T and K belong to B(X, Y ). Suppose that
T (X) is closed and K is of finite rank, then the image (T + K)(X) is also closed.
Proof. Suppose the conclusion does not hold. We write z̄ := z + ker(T + K) for z ∈ X. Then by
Corollary 9.8, there is a sequence (xn ) in X such that ∥x̄n ∥ = 1 for all n and ∥(T +K)xn ∥ → 0. Thus,
(xn ) can be chosen so that it is bounded. By passing a subsequence of (xn ) we may assume that
y := limn K(xn ) exists in Y because K is of finite rank. Therefore, we have limn T (xn ) = −y. Since
T has closed range, we have T x = −y for some x ∈ X. This gives lim T (xn − x) = 0. Note that the
natural map Te is a topological isomorphism from X/ ker T onto T (X) because T (X) is closed. We
see that ∥xn −x+ker T ∥ → 0 and thus, ∥y−K(x)+K(ker T )∥ = lim ∥K(xn )−K(x)+K(ker T )∥ = 0.
From this we have y − Kx = Ku for some u ∈ ker T . In addition, for each n, there is an element
tn ∈ ker T so that ∥xn − x + tn ∥ < 1/n. This implies that
∥K(tn − u)∥ ≤ ∥K(tn + (xn − x))∥ + ∥ − K(xn + x) − K(u)∥ ≤ ∥K∥1/n → 0.
Therefore, we have ∥tn − u + (ker T ∩ ker K)∥ → 0 because tn − u ∈ ker T and the image of K| ker T
is closed. From this we see that ∥tn − u + ker(T + K)∥ → 0.
On the other hand, since T x = −y = −Kx − Ku and u ∈ ker T , we have (T + K)x = −Ku − T u
and so, x + u ∈ ker(T + K). Then we can now conclude that
∥x̄n ∥ = ∥x̄n − (x̄ + ū)∥ ≤ ∥x̄n − x̄ − t̄n ∥ + ∥t̄n − ū∥ → 0.
It contradicts to the choice of xn such that ∥x̄n ∥ = 1 for all n. The proof is complete. □
Remark 9.10. In general, the sum of operators of closed ranges may not have a closed range.
Before looking for those examples, let us show the following simple useful lemma.
30 CHI-WAI LEUNG
Lemma 9.11. Let X be a Banach space. If T ∈ B(X) with ∥T ∥ < 1, then the operator 1 − T is
invertible, i.e., there is S ∈ B(X) such that (1 − T )S = S(1 − T ) = 1.
Proof. Note that since X is a Banach space, the set of all bounded
P operators B(X) is a Banach
space under the usual operator norm. This implies that the series ∞k=0 T k is convergent in B(X)
Xn
because ∥T ∥ < 1. On the other hand, we have 1 − T n = (1 − T )( T k ) for all n = 1, 2.... Taking
k=0
∞
X
n → ∞, we see that (1 − T )−1 exists, in fact, (1 − T )−1 = T k. □
k=0
Theorem 10.1. Closed Graph Theorem : Let T : X −→ Y be a linear operator from a Banach
space X to a Banach Y . Then T is bounded if and only if T is closed.
Proof. The part (⇒) is clear.
Assume that T is closed, i.e., the graph G(T ) is ∥ · ∥∞ -closed. Define ∥ · ∥0 : X −→ [0, ∞) by
∥x∥0 = ∥x∥ + ∥T (x)∥
for x ∈ X. Then ∥ · ∥0 is a norm on X. Let I : (X, ∥ · ∥0 ) −→ (X, ∥ · ∥) be the identity operator. It
is clear that I is bounded since ∥ · ∥ ≤ ∥ · ∥0 .
Claim: (X, ∥ · ∥0 ) is Banach. In fact, let (xn ) be a Cauchy sequence in (X, ∥ · ∥0 ). Then (xn ) and
(T (xn )) both are Cauchy sequences in (X, ∥ · ∥) and (Y, ∥ · ∥Y ). Since X and Y are Banach spaces,
there are x ∈ X and y ∈ Y such that ∥xn − x∥X → 0 and ∥T (xn ) − y∥Y → 0. Thus y = T (x) since
the graph G(T ) is closed.
Then by Theorem 9.7, the norms ∥ · ∥ and ∥ · ∥0 are equivalent. Hence, there is c > 0 such that
∥T (·)∥ ≤ ∥ · ∥0 ≤ c∥ · ∥ and hence, T is bounded since ∥T (·)∥ ≤ ∥ · ∥0 . The proof is complete. □
P∞ 2
Example 10.2. Let D := {c = (cn ) ∈ ℓ2 : 2 2
n=1 n |cn | < ∞}. Define T : D −→ ℓ by
T (c) = (ncn ). Then T is an unbounded closed operator.
31
Proof. Note that since ∥T en ∥ = n for all n, T is not bounded. Now we claim that T is closed.
Let (xi ) be a convergent sequence in D such that (T xi ) is also convergent in ℓ2 . Write xi = (xi,n )∞
n=1
with lim xi = x := (xn ) in D and lim T xi = y := (yn ) in ℓ2 . This implies that if we fix n0 , then
i i
lim xi,n0 = xn0 and lim n0 xi,n0 = yn0 . This gives n0 xn0 = yn0 . Thus T x = y and hence T is
i i
closed. □
Let X be a normed space and let X ∗ be its dual space. Then there is a natural bi-linear mapping
on X × X ∗ (call a dual pair) given by
⟨·, ·⟩ : X × X ∗ → K; ⟨x, f ⟩ = f (x).
Moreover, this dual pair is non-degenerate, that is, ⟨x, f ⟩ = 0 for all f ∈ X ∗ if and only if x = 0
and ⟨x, f ⟩ = 0 for all x ∈ X if and only if f = 0.
Proof. Note that by the assumption, we can define a linear operator T from X to Y given by
T x := limk Tk x for x ∈ X. We need to show that T is bounded. In fact, (∥Tk ∥) is bounded by the
Uniform Boundedness Theorem since limk Tk x exists for all x ∈ X. Hence, for each x ∈ BX , there
is a positive integer K such that ∥T x∥ ≤ ∥TK x∥ + 1 ≤ (supk ∥Tk ∥) + 1. Thus, T is bounded.
Finally, it remains to show the last assertion. In fact, note that for any x ∈ BX and ε > 0, there
is N (x) ∈ N such that ∥T x∥ < ∥Tk x∥ + ε < ∥Tk ∥ + ε for all k ≥ N (x). This gives ∥T x∥ ≤
inf k≥N (x) ∥Tk ∥ + ε for all k ≥ N (x) and hence, ∥T x∥ ≤ inf k≥N (x) ∥Tk ∥ + ε ≤ supn inf k≥n ∥Tk ∥ + ε
for all x ∈ BX and ε > 0. Therefore, we have ∥T ∥ ≤ lim inf ∥Tk ∥. □
k
Corollary 11.4. Every weakly convergent sequence in a normed space must be bounded.
Proof. Let (xn ) be a weakly convergent sequence in a normed space X. If we let Q : X → X ∗∗
be the canonical isometry, then (Qxn ) is a bounded sequence in X ∗∗ . Note that (xn ) is weakly
convergent if and only if (Qxn ) is w∗ -convergent. Thus, (Qxn (x∗ )) is bounded for all x∗ ∈ X ∗ .
Note that the dual space X ∗ must be complete. We can apply the Uniform Boundedness Theorem
to see that (Qxn ) is bounded and so is (xn ). □
(xn ) is a sequence in E such that lim xn = x and lim Qxn = u for some x, u ∈ E, then Qx = u.
Indeed, if we let xn = yn + zn where yn ∈ E and zn ∈ F , then Qxn = yn . Note that (zn ) is
a convergent sequence in F because zn = xn − yn and (xn ) and (yn ) both are convergent. Let
w = lim zn . This implies that
x = lim xn = lim(yn + zn ) = u + w.
Since E and F are closed, we have u ∈ E and w ∈ F . Therefore, we have Qx = u as desired.
The converse is clear. In fact, we have X = im Q ⊕ ker Q in this case. □
Recall that a closed subspace M of a Banach space E is called an M -ideal if the space M ⊥ :=
{x∗ ∈ E ∗ : x∗ (M ) ≡ 0} is L
a ℓ1 -direct summand of E ∗ , that is, there is another closed subspace N
∗ ∗ ⊥ ∗ ∗ ∗
of E such that E = M ℓ1 N , i.e., for every element x ∈ E satisfies the condition: x = u + v
and ∥x∗ ∥ = ∥u∥ + ∥v∥ for a pair of elements u and v in M ⊥ and N respectively.
n
Proposition 12.7. The c0 space is an M -ideal of ℓ∞ .
Proof. We first notice that for h ∈ (ℓ∞ )∗ and ξ ∈ ℓ∞ , then Re(h)(ξ) := Re(h(ξ)) can be viewed as
a R-linear functional on ℓ∞ and ∥h∥ = ∥Re(h)∥.
Using Proposition 12.6, it suffices to show that for g ∈ c∗0 = ℓ1 and f ∈ c⊥
0 , we have ∥g + f ∥(ℓ∞ )∗ =
⊥ ∗
∥g∥(ℓ∞ )∗ + ∥f ∥(ℓ∞ )∗ , where c0 := {f ∈ (ℓ∞ ) : f (c0 ) ≡ 0}. Let ε > 0. By considering the polar
35
decomposition, then there are elements ξ and ξ ′ in (ℓ∞ )1 of norm-one such that
∞
X
∥f ∥ − ε < f (ξ) and ∥g∥ − ε < g(ξ ′ ) = Re(g)(ξ ′ ) = Re(ξ ′ (n)g(n)).
n=1
Since g ∈ c∗0 = ℓ1 , there is N such that |g(n)| < ε. Now let ξ ′′ be an element in ℓ∞ given by
P
n>N
(
ξ ′ (n) if n ≤ N
ξ ′′ (n) :=
ξ(n) if n > N .
Then ∥ξ ′′ ∥∞ ≤ 1 and ξ ′′ −P
ξ ∈ c00 . Hence we have f (ξ) = f (ξ ′′ ) because f (c0 ) ≡ 0.
On the other hand, since n>N |g(n)| < ε, we have
N
X X N
X
′ ′
∥g∥ − ε < Re(g)(ξ ) ≤ Re(ξ (n)g(n)) + |g(n)| < Re(ξ ′ (n)g(n)) + ε.
n=1 n>N n=1
Thus, we have
∞
X
′′
Re(g)(ξ ) = Re(ξ ′′ (n)g(n))
n=1
XN X
≥ Re(ξ ′ (n)g(n)) − | ξ(n)g(n)|
n=1 n>N
≥ ∥g∥ − 2ε − ε.
Therefore, we have
∥g∥ + ∥f ∥ = ∥Re(g)∥ + ∥f ||
≤ Re(g)(ξ ′′ ) + f (ξ ′′ ) + 4ε
= Re(g + f )(ξ ′′ ) + 4ε
≤ ∥Re(g + f )∥ + 4ε
= ∥g + f ∥ + 4ε.
for all ε > 0. The proof is complete.
□
Theorem 13.1. Using the notations as above, for each element x ∈ X0 , put
q(x) := sup{∥Qn (x)∥ : n = 1, 2...}.
Then
(i) q is a Banach equivalent norm on X0 .
(ii) Each coordinate projection Qn and coordinate function ψn are bounded in the original norm-
topology.
36 CHI-WAI LEUNG
Proof. Since x = limn Qn x for all x ∈ X0 , ∥x∥ = limn ∥Qn x∥. This implies that the sequence
(∥Qn x)∥) is bounded and hence, q is well defined, in addition, q(·) ≥ ∥ · ∥ on X0 . From this,
together with the Open Mapping Theorem, all assertions follows if we show that q is a Banach
norm on X0 .
Let (xn ) be a Cauchy sequence in X0 with respect to the norm q. Clearly, (xn ) is also a Cauchy
sequence in the ∥ · ∥-topology because q(·) ≥ ∥ · ∥. Let x = limn xn be the limit in X0 in the
∥ · ∥-topology. We are going to show that x is also the limit of (xn ) with respect to the q-topology.
We first note that yk := limn Qk xn exists in X0 for all k = 1, 2, ... by the definition of the norm q.
Claim 1: ∥ · ∥-limk yk = x.
Let ε > 0. Then by the definition of the norm q, there is a positive integer N1 such that ∥Qk xN −
Qk xm ∥ < ε and ∥xN − xm ∥ < ε for all m, N ≥ N1 and for all k = 1, 2.... This gives
∥x − Qk xm ∥ ≤ ∥x − xN1 ∥ + ∥xN1 − Qk xN1 ∥ + ∥Qk xN1 − Qk xm ∥ < 2ε + ∥xN1 − Qk xN1 ∥
for all m ≥ N1 and for all positive integers k. Thus, if we take m → ∞, then we have
∥x − yk ∥ ≤ 2ε + ∥xN1 − Qk xN1 ∥ → 2ε + 0 as k → ∞.
Claim 2: Qk x = yk for all k = 1, 2....
Fix a positive integer k1 . Note that Qk1 yk = yk1 for all k ≥ k1 . Indeed, since Ek and Ek1 are of
finite dimension, the restrictions Qk1 |Ek and Qk |Ek1 both are continuous. This implies that
Qk1 yk = Qk1 (lim Qk xn ) = lim Qk1 Qk (xn ) = lim Qk Qk1 (xn ) = Qk (lim Qk1 xn ) = Qk (yk1 ) = yk1
n n n n
Proposition 13.2. Using the notation as above, a sequence (xi ) is a basic sequence in X if and
only if there is K > 0 such that
X X
ai xi ≤ K ai xi ,
1≤i≤n 1≤i≤m
We first recall the following useful properties of an inner product space which can be found in the
standard text books of linear algebras.
Proposition 14.1. Let V be an inner product space. For all x, y ∈ V , we always have:
(i): (Cauchy-Schwarz inequality): |(x, y)| ≤ ∥x∥∥y∥ Consequently, the inner product on
V × V is jointly continuous.
(ii): (Parallelogram law): ∥x + y∥2 + ∥x − y∥2 = 2∥x∥2 + 2∥y∥2 .
Furthermore, a norm ∥ · ∥ on a vector space X is induced by an inner product if and only if
it satisfies the Parallelogram law. In this case such inner product is given by the following:
1 1
Re(x, y) = (∥x + y∥2 − ∥x − y∥2 ) and Im(x, y) = (∥x + iy∥2 − ∥x − iy∥2 )
4 4
for all x, y ∈ X.
(iii) Gram-Schmidt process Let {x1 , x2 , ...} be a sequence of linearly independent vectors in
an inner product space V . Put e1 := x1 /∥x1 ∥. Define en inductively on n by
xn − nk=1 (xk , ek )ek
P
en+1 := .
∥xn − nk=1 (xk , ek )ek ∥
P
Then {en : n = 1, 2, ..} forms an orthonormal system in V Moreover, the linear span of
x1 , ..., xn is equal to the linear span of e1 , ..., en for all n = 1, 2....
Definition 14.3. A Hilbert space is a Banach space whose norm is given by an inner product.
Example 14.4. (1) CN is a Hilbert space under the usual inner product given by (w, z) :=
PN N
k=1 w(k)z(k) for w, z ∈ C .
(2) It follows from Proposition 14.1 immediately that ℓ2 is a Hilbert space and ℓp is not a Hilbert
space for all p ∈ [1, ∞] \ {2}.
38 CHI-WAI LEUNG
In the rest of this section, X always denotes a complex Hilbert space with an inner product (·, ·).
Recall that two vectors x and y in an inner product space V are said to be orthogonal if (x, y) = 0.
Proposition 14.5. (Bessel′ s inequality) : Let {e1 , ..., eN } be an orthonormal set in an inner
product space V , i.e., (ei , ej ) = 1 if i = j, otherwise is equal to 0. Then for any x ∈ V , we have
N
X
|(x, ei )|2 ≤ ∥x∥2 .
i=1
Corollary 14.6. Let (ei )i∈I be an orthonormal set in a Hilbert space X. Then for any element
x ∈ X, the set
I(x) := {i ∈ I : (ei , x) ̸= 0}
is countable.
Proof. Note that for each x ∈ V , we have
∞
[
{i ∈ I : (ei , x) ̸= 0} = {i ∈ I : |(ei , x)| ≥ 1/n}.
n=1
Then the Bessel’s inequality implies that the set {i ∈ I : |(ei , x)| ≥ 1/n} must be finite for each
n ≥ 1. Thus the result follows. □
Proposition 14.7. Let P (en ) be a sequence of orthonormal vectors in a Hilbert space X. Then for
any x ∈ V , the series ∞ n=1 (x, en )en is convergent.
Moreover, if (eσ(n) ) is a rearrangement of (en ), i.e., σ : {1, 2...} −→ {1, 2, ..} is a bijection. Then
we have
X∞ ∞
X
(x, en )en = (x, eσ(n) )eσ(n) .
n=1 n=1
of the series ∞
P
Proof. Since X is a Hilbert space, the convergence
Pp n=1 (x, en )en follows from the
Bessel’s inequality. In fact, if we put sp := n=1 (x, en )en , then we have
X
∥sp+k − sp ∥2 = |(x, en )|2 .
p+1≤n≤p+k
P∞ P∞
Now put y = n=1 (x, en )en and z = n=1 (x, eσ(n) )eσ(n) . Note that we have
N
X N
X
(y, y − z) = lim( (x, en )en , (x, en )en − z)
N
n=1 n=1
N
X N
X ∞
X
2
= lim |(x, en )| − lim (x, en ) (x, eσ(j) )(en , eσ(j) )
N N
n=1 n=1 j=1
∞
X N
X
2
= |(x, en )| − lim (x, en )(x, en ) (∵ for each n, there is a unique j such that n = σ(j))
N
n=1 n=1
= 0.
Similarly, we have (z, y − z) = 0. The result follows. □
39
Proposition 14.8. Let {ei }i∈I be a family of orthonormal vectors in X. Then the followings are
equivalent:
(i): {ei }i∈I is complete;
(ii): if (x, ei ) = 0 for all i ∈ I, then
P x = 0;
(iii): for any x ∈ X, we have x = i∈I (x, ei )ei ;
(iv): for any x ∈ X, we have ∥x∥2 = i∈I |(x, ei )|2 .
P
Note : there are only countable many (x, ei ) ̸= 0 by Corollary 14.6, so the sums in (iii) and (iv)
are convergent by Proposition 14.7. In this case, the expression of each element x ∈ X in Part (iii)
is unique.
Remark 14.10. Recall that a vector space dimension of X is defined by the cardinality of a maximal
linearly independent set in X.
Note that if X is finite dimensional, then the orthonormal dimension is the same as the vector
space dimension.
In addition, the vector space dimension is larger than the orthornormal dimension in general since
every orthogonal set must be linearly independent.
Two Hilbert spaces X and Y are said to be isomorphic if there is linear isomorphism U from X
onto Y such that (U x, U x′ ) = (x, x′ ) for all x, x′ ∈ X. In this case U is called a unitary operator.
Theorem 14.11. Two Hilbert spaces are isomorphic if and only if they have the same orthonornmal
dimension.
Proof. The converse part (⇐) is clear.
Now for the (⇒) part, let X and Y be isomorphic Hilbert spaces. Let U : X −→ Y be a unitary.
Note that if {ei }i∈I is an orthonormal basis of X, then {U ei }i∈I is also an orthonormal basis of Y .
Thus the necessary part follows immediately from Proposition 14.9. □
40 CHI-WAI LEUNG
for each f, g ∈ C(T). We write ∥ · ∥2 for the norm induced by this inner product.
The Hilbert space L2 (T) is defined by the completion of C(T) under the norm ∥ · ∥2 .
Now for each n ∈ Z, put fn (z) = z n . We claim that {fn : n = 0, ±1, ±2, ....} is an orthonor-
mal basis for L2 (T).
In fact, by using the Euler Formula: eiθ = cos θ + i sin θ for θ ∈ R, we see that the family
{fn : n ∈ Z} is orthonormal.
It remains to show that the family {fn } is maximal. By Proposition 14.8, it needs to show
that if (g, fn ) = 0 for all n ∈ Z, then g = 0 in L2 (T). for showing this, we have to make
use the known fact that every element in L2 (T) can be approximated by the polynomial
functions of z and z̄ on T in ∥ · ∥2 -norm due to the the Stone-Weierstrass Theorem:
For a compact metric space E, suppose that a complex subalgebra A of C(E) satisfies the
conditions: (i): the conjugate f¯ ∈ A whenever f ∈ A, (ii): for every pair z, z ′ ∈ E, there is
f ∈ A such that f (z) ̸= f (z ′ ) and (iii): A contains the constant one function. Then A is
dense in C(E) with respect to the sup-norm.
Thus, the algebra of all polynomials functions of z and z̄ on T is dense in C(T). From
this we can find a sequence of polynomials (pn (z, z̄)) such that ∥g − pn ∥2 → 0 as n → 0.
Since (g, fn ) = 0 for all n, we see that (g, pn ) = 0 for all n. Therefore, we have
∥g∥22 = lim(g, pn ) = 0.
n
Remark 14.14. In view of Example 14.13(iii) above, L2 (T) can be identified as the space
R 2π
L2 [0, 2π] := {f : [0, 2π] → C : f is a Lebesgue measure function so that 0 |f (x)|2 dx exists}.
1
R 2π
In this case, the inner product in L2 [0, 2π] is given by (f, g) := 2π 0 f (t)g(t)dt. Now for each
x ∈ [0, 2π] and f ∈ L2 [0, 2π], put ek (x) := eikx and
N N Z 2π
X 1 X
(14.1) SN (f, x) := (f, ek )ek (x) = ( f (t)e−ik dt)eikx .
2π 0
k=−N k=−N
∞
X
Then the series (f, ek )ek (x) (need not be convergent) is the usual notation of the Fourier series
−∞
of f at x.
N
X
In the previous result, we have seen that the series lim ∥f − (f, ek )ek ∥L2 → 0
N →∞
k=−N
It is naturally raised the question: for f ∈ C[0, 2π], do we have f (x) = lim SN (f, x) for all
N →∞
x ∈ [0, 2π], that is, f can be written as its Fourier series? The answer is negative. Indeed, we have
the following stronger result.
Theorem 14.15. For each element x0 ∈ [0, 2π], there is a continuous function f on [0, 2π] such
∞
X
that its Fourier series (f, ek )eikx0 at x0 is divergent.
k=−∞
Proof. We give an outline argument in here. The details of the proof is referred to the Katznelson’s
classic book [9].
We keep the notation as above. We fix a point x0 ∈ [0, 2π]. For f ∈ C[0, 2π], Using the Eq 14.1
above, we have
N Z 2π
1 X
φN (f ) := SN (f, x0 ) = ( f (t)e−ik dt)eikx0
2π 0
k=−N
Z 2π N
(14.2) 1 X
= f (t)( eik(x0 −t) )dt
2π 0 k=−N
Z 2π
1
= f (t)DN (x0 − t)dt,
2π 0
N
X
where DN (s) := eiks for s ∈ R. Then φN ∈ C[0, 2π]∗ . It is a fact that we have (see [9,
k=−N
Chapter 2. §2])
Z 2π
1
∥φN ∥ = |DN (x0 − t)|dt → ∞ as N → ∞.
2π 0
From this we have sup{∥φN ∥ : N = 1, 2....} = ∞. This implies that the following set is dense in
C[0, 2π].
D := {f ∈ C[0, 2π] : sup |φN (f )| = ∞}.
N
To see this, suppose not. Then there is an element f0 ∈ C[0, 2π] and r > 0 such that supN |φN (f )| <
∞ for all f ∈ B(f0 , r). By considering B(0, 1) = 1r (B(f0 , r) − f0 ), then we have supN |φN (g)| < ∞
for all g ∈ B(0, 1) and so, supN |φN (g)| < ∞ for all g ∈ C[0, 2π]. The Uniform Boundedness
Theorem implies that supN ∥φN ∥ < ∞ that leads to a contradiction.
42 CHI-WAI LEUNG
∞
X
On the other hand, it is clear that if the series (f, ek )eikx0 is convergent, then supN |SN (f, x0 )| <
k=−∞
∞. Therefore, the Fourier series of each element in D is divergent at x0 as desired. We finish the
proof. □
Students should be made to think, to doubt, to communicate, to question, to learn from their
mistakes, and most importantly have fun in their learning.
∼ Richard Feynman ∼ one of the greatest physicists in 20th century.
The last assertion follows by considering the closed convex set u−D := {u−x : x ∈ D} immediately.
□
Remark 15.2. Using the notation given as in Proposition 15.1, we have a well defined function
r : X → X given by x ∈ X 7→ r(x) ∈ D such that ∥x − r(x)∥ = dist(x, D). Clearly, we have
r(x) = x whenever x ∈ D. Moreover, we have the following assertion which are shown in [5].
This gives t2 ∥x − r(z)∥2 + 2tRe(x − r(z), r(z) − z) ≥ 0 for all 0 ≤ t ≤ 1. This implies that Re(x −
r(z), r(z)−z) ≥ 0 for all x ∈ D and z ∈ X. From this, for a, b ∈ X we have Re(r(b)−r(a), r(a)−a) ≥
0 and Re(r(a)−r(b), r(b)−b) ≥ 0, so we have Re(r(b)−r(a), r(a)−a)+Re(r(b)−r(a), b−r(b)) ≥ 0.
Thus, we have
∥r(b) − r(a)∥2 = Re(r(b) − r(a), r(b) − r(a))
≤ Re(r(b) − r(a), b − a)
≤ |(r(b) − r(a), b − a)|
≤ ∥r(b) − r(a)∥∥b − a∥.
The proof is complete. □
Proposition 15.4. Suppose that M is a closed subspace. Let u ∈ X and w ∈ M . Then the
followings are equivalent:
(i): ∥u − w∥ = d(u, M );
(ii): u − w ⊥ M , i.e., (u − w, x) = 0 for all x ∈ M .
Consequently, for each element u ∈ X, there is a unique element w ∈ M such that u − w ⊥ M .
Proof. Let d := d(u, M ).
For proving (i) ⇒ (ii), fix an element x ∈ M . Then for any t > 0, note that since w + tx ∈ M , we
have
d2 ≤ ∥u − w − tx∥2 = ∥u − w∥2 + ∥tx∥2 − 2Re(u − w, tx) = d2 + ∥tx∥2 − 2Re(u − w, tx).
This implies that
(15.1) 2Re(u − w, x) ≤ t∥x∥2
for all t > 0 and for all x ∈ M . Thus by considering −x in Eq.15.1, we obtain
2|Re(u − w, x)| ≤ t∥x∥2 .
for all t > 0. This implies that Re(u − w, x) = 0 for all x ∈ M . Similarly, putting ±ix into Eq.15.1,
we have Im(u − w, x) = 0. Thus(ii) follows.
For (ii) ⇒ (i), we need to show that ∥u − w∥2 ≤ ∥u − x∥2 for all x ∈ M . Note that since u − w ⊥ M
and w ∈ M , we have u − w ⊥ w − x for all x ∈ M . This gives
∥u − x∥2 = ∥(u − w) + (w − x)∥2 = ∥u − w∥2 + ∥w − x∥2 ≥ ∥u − w∥2 .
Part (i) follows.
The last statement is obtained immediately by Proposition 15.1. □
Corollary 15.6. Let M be a closed subspace of X. Then M ⊊ X if and only if there is a non-zero
element z ∈ X such that z ⊥ M .
Proof. It is clear from Theorem 15.5. □
44 CHI-WAI LEUNG
Theorem 15.8. Riesz Representation Theorem : For each f ∈ X ∗ , then there is a unique
element vf ∈ X such that
f (x) = (x, vf )
for all x ∈ X and we have ∥f ∥ = ∥vf ∥.
P
Furthermore, if (ei )i∈I is an orthonormal basis of X, then vf = i f (ei )ei .
Proof. We first prove the uniqueness of vf . If z ∈ X also satisfies the condition: f (x) = (x, z) for
all x ∈ X. This implies that (x, z − vf ) = 0 for all x ∈ X. Thusz − vf = 0.
Now for proving the existence of vf , it suffices to show the case ∥f ∥ = 1. Then ker f is a closed
proper subspace. Then by the orthogonal decomposition again, we have
X = ker f ⊕ (ker f )⊥ .
Since f ̸= 0, we have (ker f )⊥ is linear isomorphic to C. Note that the restriction of f on (ker f )⊥ is
of norm one. Hence there is an element vf ∈ (ker f )⊥ with ∥vf ∥ = 1 such that f (vf ) = ∥f |(ker f )⊥ ∥ =
1 and (ker f )⊥ = Cvf . Thusfor each element x ∈ X, we have x = z + αvf for some z ∈ ker f and
α ∈ C. Then f (x) = αf (vf ) = α = (x, vf ) for all x ∈ X.X
Concerning about the last assertion, if we put vf = αi ei , then f (ej ) = (ej , vf ) = αj for all
i∈I
j ∈ I. □
2 ∗
Example R 15.9. Consider the Hilbert space H := L (T) (see Example 14.13). Define φ ∈ H by
R element g ∈ H, there is an element h ∈ ker φ
φ(f ) := T f (z)dz. Using Proposition 15.4, for each
such that ∥g −h∥ = dist(g, ker φ). Then h = g −( gdz)1 where 1 denotes the constant-one function
on T. In fact, consider the orthogonal decomposition H = ker φ ⊕ (ker φ)⊥ . Note that φ(g) = (g, 1)
for all g ∈R H. Thus, for each g ∈ H, we have g = h ⊕ α1. From this, we see that α = (g, 1). Thus,
h = g − ( gdz)1.
Corollary 15.10. Using the notations as in Theorem 15.8, define the map
(15.2) Φ : f ∈ X ∗ 7→ vf ∈ X, i.e., f (y) = (x, Φ(f ))
for all y ∈ X and f ∈ X ∗ .
Moreover, if we define (f, g)X ∗ := (vg , vf )X for f, g ∈ X ∗ , then (X ∗ , (·, ·)X ∗ ) becomes a Hilbert
space, and Φ is an anti-unitary operator from X ∗ onto X, i.e., Φ satisfies the conditions:
Φ(αf + βg) = αΦ(f ) + βΦ(g) and (Φf, Φg)X = (g, f )X ∗
for all f, g ∈ X ∗ and α, β ∈ C.
Furthermore, if we define J : x ∈ X 7→ fx ∈ X ∗ , where fx (y) := (y, x), then J is the inverse of Φ,
and hence, J is an isometric conjugate linear isomorphism.
Proof. The result follows immediately from the observation that vf +g = vf + vg and vαf = αvf for
all f ∈ X ∗ and α ∈ C.
The last assertion is clearly obtained by the Eq.15.2 above. □
Proof. Using the notations as in the Riesz Representation Theorem 15.8, let X be a Hilbert space.
and Q : X → X ∗∗ the canonical isometry. Let ψ ∈ X ∗∗ . To apply the Riesz Theorem on the dual
space X ∗ , there exists an element x∗0 ∈ X ∗ such that
ψ(f ) = (f, x∗0 )X ∗
for all f ∈ X ∗ . By using Corollary 15.10, there is an element x0 ∈ X such that x0 = vx∗0 and thus,
we have
ψ(f ) = (f, x∗0 )X ∗ = (x0 , vf )X = f (x0 )
for all f ∈ X ∗ . Therefore, ψ = Q(x0 ) and so, X is reflexive.
The proof is complete. □
Theorem 15.12. Every bounded sequence in a Hilbert space has a weakly convergent subsequence.
Proof. Let (xn ) be a bounded sequence in a Hilbert space X and M be the closed subspace of X
spanned by {xm : m = 1, 2...}. Then M is a separable Hilbert space.
Method I : Define a map by jM : x ∈ M 7→ jM (x) := (·, x) ∈ M ∗ . Then (jM (xn )) is a bounded
sequence in M ∗ . By Banach’s result, Proposition 6.9, (jM (xn )) has a w∗ -convergent subsequence
w∗
(jM (xnk )). Put jM (xnk ) −−→ f ∈ M ∗ , i.e., jM (xnk )(z) → f (z) for all z ∈ M . The Riesz Rep-
resentation will assure that there is a unique element m ∈ M such that jM (m) = f . Thuswe
have (z, xnk ) → (z, m) for all z ∈ M . In particular, if we consider the orthogonal decomposition
X = M ⊕ M ⊥ , then (x, xnk ) → (x, m) for all x ∈ X and thus (xnk , x) → (m, x) for all x ∈ X. Then
xnk → m weakly in X by using the Riesz Representation Theorem again.
Method II : We first note that since M is a separable Hilbert space, the second dual M ∗∗ is also
separable by the reflexivity of M . Thus, the dual space M ∗ is separable (see Proposition4.11).
Let Q : M −→ M ∗∗ be the natural canonical mapping. To apply the Banach’s result Propo-
sition 6.9 for X ∗ , then Q(xn ) has a w∗ -convergent subsequence, says Q(xnk ). This gives an
element m ∈ M such that Q(m) = w∗ -limk Q(xnk ) because M is reflexive. Thus, we have
f (xnk ) = Q(xnk )(f ) → Q(m)(f ) = f (m) for all f ∈ M ∗ . Using the same argument as in Method I
again, xnk weakly converges to m. □
Remark 15.13. Recall the well known James’s Theorem that a Banach space X is reflexive if
and only if every bounded sequence in X has a weakly convergent subsequence. (see Appendix 7
for the proof of a separable space case). Hence, Theorem 15.12 can be obtained by the James’s
Theorem directly. However, the Riesz Representation Theorem gives a simple proof for the Hilbert
spaces case.
Remark 16.1. For Hilbert spaces H1 and H2 , we consider their direct sum H := H1 ⊕ H2 . If we
define the inner product on H by
(x1 ⊕ x2 , y1 ⊕ y2 ) := (x1 , y1 )H1 + (x2 , y2 )H2
46 CHI-WAI LEUNG
for x1 ⊕ x2 and y1 ⊕ y2 in H, then H becomes a Hilbert space. Now for each T ∈ B(H1 , H2 ), we
can define an element T̃ ∈ B(H) by T̃ (x1 ⊕ x2 ) := 0 ⊕ T x1 . Therefore, the space B(H1 , H2 ) can be
viewed as a closed subspace of B(H). Thus, we can consider the case of H1 = H2 for studying the
space B(H1 , H2 ).
Proposition 16.3. Let T ∈ B(X). Then there is a unique element T ∗ in B(X) such that
(16.2) (T x, y) = (x, T ∗ y)
In this case, T ∗ is called the adjoint operator of T .
Proof. First, we show the uniqueness. Suppose that there are S1 , S2 in B(X) which satisfy the
Eq.16.2. Then (x, S1 y) = (x, S2 y) for all x, y ∈ X. Eq.16.1 implies that S1 = S2 .
Finally, we prove the existence. Note that if we fix an element y ∈ X, define the map fy (x) :=
(T x, y) for all x ∈ X. Then fy ∈ X ∗ . By applying the Riesz Representation Theorem, there is a
unique element y ∗ ∈ X such that (T x, y) = (x, y ∗ ) for all x ∈ X and ∥fy ∥ = ∥y ∗ ∥. In addition, we
have
|fy (x)| = |(T x, y)| ≤ ∥T ∥∥x∥∥y∥
for all x, y ∈ X and thus ∥fy ∥ ≤ ∥T ∥∥y∥. If we put T ∗ (y) := y ∗ , then T ∗ satisfies the Eq.16.2.
Moreover, we have ∥T ∗ y∥ = ∥y ∗ ∥ = ∥fy ∥ ≤ ∥T ∥∥y∥ for all y ∈ X. Thus, we have T ∗ ∈ B(X) and
∥T ∗ ∥ ≤ ∥T ∥. Hence the operator T ∗ is as desired. □
Example 16.7. If X = Cn and D = (aij )n×n an n × n matrix, then D∗ = (aji )n×n . In fact, note
that
aji = (Dei , ej ) = (ei , D∗ ej ) = (D∗ ej , ei ).
Thusif we put D∗ = (dij )n×n , then dij = (D∗ ej , ei ) = aji .
∞
P∞ X
Example 16.8. Let ℓ2 (N) := {x : N → C : 2
i=0 |x(i)| < ∞}, and put (x, y) := x(i)y(i).
i=0
Define the operator D ∈ B(ℓ2 (N)) (called the unilateral shift) by
Dx(i) = x(i − 1)
for i ∈ N, where we set x(−1) := 0, i.e., D(x(0), x(1), ...) = (0, x(0), x(1), ....).
Then D is an isometry and the adjoint operator D∗ is given by
D∗ x(i) := x(i + 1)
for i = 0, 1, .., i.e., D∗ (x(0), x(1), ...) = (x(1), x(2), ....).
Indeed we can directly check that
∞
X ∞
X
(Dx, y) = x(i − 1)y(i) = x(j)y(j + 1) = (x, D∗ y).
i=0 j=0
Example 16.9. Let ℓ∞ (N) = {x : N → C : supi≥0 |x(i)| < ∞} and ∥x∥∞ := supi≥0 |x(i)|. For
each x ∈ ℓ∞ , define Mx ∈ B(ℓ2 (N)) by
Mx (ξ) := x · ξ
for ξ ∈ ℓ2 (N),
where (x · ξ)(i) := x(i)ξ(i); i ∈ N.
Then ∥Mx ∥ = ∥x∥∞ and Mx∗ = Mx , where x(i) := x(i).
Definition 16.10. Let T ∈ B(X) and let I be the identity operator on X. T is said to be
(i) : selfadjoint if T ∗ = T ;
(ii) : normal if T ∗ T = T T ∗ ;
(iii) : unitary if T ∗ T = T T ∗ = I.
Remark 16.12. In Proposition 16.11(i), if the domain of T is replaced P∞ by dense domain, then
the conclusion does not hold. For example, let D := {x ∈ ℓ2 : n=1 |nx(n)|2 < ∞} and let
T (x)(n) := nx(n) for x ∈ D. Then D is a dense domain because the canonical basis (en ) ⊆ D. It
is noted that T is unbounded on D, but (T x, y) = (x, T y) for all x, y ∈ D.
Proposition 16.15. Let X be a Hilbert space. Let M and N be the closed subspaces of X such
that
X =M ⊕N . . . . . . . . . . . . (∗)
Let Q : X → X be the projection along the decomposition (∗) with im Q = M (note that Q is
bounded by Proposition 12.1). Then N = M ⊥ (and hence (∗) is the orthogonal decomposition of X
with respect to M ) if and only if Q satisfies the conditions: Q2 = Q and Q∗ = Q. In this case, Q
is called the orthogonal projection (or projection for simply) with respect to M .
Proof. Now if N = M ⊥ , then for y, y ′ ∈ M and z, z ′ ∈ N , we have
(Q(y + z), y ′ + z ′ ) = (y, y ′ ) = (y + z, Q(y ′ + z ′ )).
ThusQ∗ = Q.
The converse of the last statement follows immediately from Proposition 16.14 because ker Q = N
and imQ = M .
The proof is complete. □
Proposition 16.16. When X is a Hilbert space, we put M the set of all closed subspaces of X and
P the set of all orthogonal projections on X. Now for each M ∈ M, let PM be the corresponding
projection with respect to the orthogonal decomposition X = M ⊕ M ⊥ . Then there is an one-one
correspondence between M and P which is defined by
M ∈ M 7→ PM ∈ P.
Furthermore, if M, N ∈ M, then we have
(i) : M ⊆ N if and only if PM PN = PN PM = PM .
(ii) : M ⊥N if and only if PM PN = PN PM = 0.
Proof. Using Proposition 16.15, we note that PM ∈ P.
Indeed the inverse of the correspondence is given by the following. If we let Q ∈ P and M = Q(X),
then M is closed. In addition, clearly we have X = Q(X) ⊕ (I − Q)X with M ⊥ = (I − Q)X. Hence
M is the corresponding closed subspace of X, i.e., M ∈ M and PM = Q.
For the final assertion, Part (i) and (ii) follow immediately from the orthogonal decompositions
X = M ⊕ M ⊥ = N ⊕ N ⊥ and together with the fact that M ⊆ N if and only if N ⊥ ⊆ M ⊥ . □
Remark 17.2. More precisely, for a normed space E, an operator T ∈ B(E) is said to be invertible
in B(E) if T is an linear isomorphism and the inverse T −1 is also bounded. However, if E is
complete, the Open Mapping Theorem assures that the inverse T −1 is bounded automatically. Thus
if E is a Banach space and T ∈ B(E), then λ ∈ / σ(T ) if and only if T − λ := T − λI is an linear
isomorphism. Thus, λ lies in the spectrum σ(T ) if and only if T − λ is either not one-one or not
surjective.
In particular, if there is a non-zero element v ∈ X such that T v = λv, then λ ∈ σ(T ) and λ is
called an eigenvalue of T with eigenvector v.
In addition, we write σp (T ) for the set of all eigenvalue of T and call σp (T ) the point spectrum.
Example 17.3. Let E = Cn and T = (aij )n×n ∈ Mn (C). Then λ ∈ σ(T ) if and only if λ is an
eigenvalue of T and thus σ(T ) = σp (T ).
Example 17.4. Let E = (c00 (N), ∥ · ∥∞ ) (note that c00 (N) is not a Banach space). Define the
map T : c00 (N) → c00 (N) by
x(k)
T x(k) :=
k+1
for x ∈ c00 (N) and i ∈ N.
Then T is bounded, in fact, ∥T x∥∞ ≤ ∥x∥∞ for all x ∈ c00 (N).
On the other hand, we note that if λ ∈ C and x ∈ c00 (N), then
1
(T − λ)x(k) = ( − λ)x(k).
k+1
From this we see that σp (T ) = {1, 21 , 31 , ...}. In addition, if λ ∈ / {1, 12 , 13 , ...}, then T − λ is an linear
isomorphism and its inverse is given by
1
(T − λ)−1 x(k) = ( − λ)−1 x(k).
k+1
Thus, (T − λ)−1 is unbounded if λ = 0,so 0 ∈ σ(T ).
Besides, if λ ∈/ {0, 1, 21 , 31 , ...}, then (T − λ)−1 is bounded. In fact, if λ = a + ib ̸= 0, for a, b ∈ R,
1
then η := min | − a|2 + |b|2 > 0 because λ ∈ / {1, 12 , 13 , ...}. This gives
k 1+k
1
∥(T − λ)−1 ∥ = sup |( − λ)−1 | < η −1 < ∞.
k∈N k + 1
Let c ∈ C \ σ(T ). We need to find r > 0 such that µ ∈ / σ(T ) as |µ − c| < r. Note that since T − c
is invertible, then for µ ∈ C, we have T − µ = (T − c) − (µ − c) = (T − c)(I − (µ − c)(T − c)−1 ).
Therefore, if ∥(µ − c)(T − c)−1 )∥ < 1, then T − µ is invertible by Part (i). Thus, if we take
1
0<r< , then r is as desired, i.e., B(c, r) ⊆ C \ σ(T ). Hence σ(T ) is closed.
∥(T − c)−1 ∥
For the last assertion, let T ∈ GL(E). Note that for any S ∈ B(E), we have S = S − T + T =
T (1 − T −1 (T − S)). Thus, if 1 − T −1 (T − S) is invertible, then so is S. Using Part (i), if
∥T − S∥ < 1/∥T −1 ∥, then 1 − T −1 (T − S) is invertible. Therefore we have B(T, ∥T 1−1 ∥ ) ⊆ GL(E).
Finally, we show the inverse map is continuous. It suffices to show that if (Tn ) is Pa sequence in
GL(E) so that Tn → I, then Tn−1 → 1. Note that if ∥Tn − 1∥ < 1/2, then Tn−1 = ∞ k
k=0 (1 − Tn ) ,
hence, we may assume that (Tn−1 ) is uniformly bounded by 2. Therefore,
∥Tn−1 − 1∥ ≤ ∥Tn−1 ∥∥Tn − 1∥ ≤ 2∥Tn − 1∥.
The proof is complete. □
Corollary 17.6. If U is a unitary operator on a Hilbert space X, then σ(U ) ⊆ {λ ∈ C : |λ| = 1}.
Proof. Since ∥U ∥ = 1, we have σ(U ) ⊆ {λ ∈ C : |λ| ≤ 1} by Proposition 17.5(ii).
Now if |λ| < 1, then ∥λU ∗ ∥ < 1. By using Proposition 17.5 again, we have I − λU ∗ is invertible.
This implies that U − λ = U (I − λU ∗ ) is invertible and thus λ ∈
/ σ(U ). □
Example 17.7. Let E = ℓ2 (N) and let D ∈ B(E) be the right unilateral shift operator as in
Example16.8. Recall that Dx(k) := x(k − 1) for k ∈ N and x(−1) := 0. Then σp (D) = ∅ and
σ(D) = {λ ∈ C : |λ| ≤ 1}.
We first claim that σp (D) = ∅.
Suppose that λ ∈ C and x ∈ ℓ2 (N) satisfy the equation Dx = λx. Then by the definition of D, we
have
x(k − 1) = λx(k) · · · · · · · · · (∗)
for all k ∈ N.
If λ ̸= 0, then we have x(k) = λ−1 xk−1 for all k ∈ N. Since x(−1) = 0, this forces x(k) = 0 for all
i, i.e., x = 0 in ℓ2 (N).
On the other hand if λ = 0, the Eq.(∗) gives x(k − 1) = 0 for all k and so x = 0 again.
Therefore σp (D) = ∅.
Finally, we are going to show σ(D) = {λ ∈ C : |λ| ≤ 1}.
Note that since D is an isometry, ∥D∥ = 1. Proposition 17.5 tells us that
σ(D) ⊆ {λ ∈ C : |λ| ≤ 1}.
Note that since σp (D) is empty, it suffices to show that D − µ is not surjective for all µ ∈ C with
|µ| ≤ 1.
Now suppose that there is λ ∈ C with |λ| ≤ 1 such that D − λ is surjective.
We consider the case where |λ| = 1 first.
Let e1 = (1, 0, 0, ...) ∈ ℓ2 (N). Then by the assumption, there is x ∈ ℓ2 (N) such that (D − λ)x = e1
and thus Dx = λx + e1 . This implies that
x(k − 1) = Dx(k) = λx(k) + e1 (k)
for all k ∈ N. From this we have x(0) = −λ−1 and x(k) = −λ−k x(0) for all k ≥ 1 because e1 (0) = 1
and e1 (k) = 0 for all k ≥ 1. Moreover, since |λ| = 1, it turns out that |x(0)| = |x(k)| for all k ≥ 1.
As x ∈ ℓ2 (N), this forces x = 0. However, it is absurd because Dx = λx + e1 .
Now we consider the case where |λ| < 1.
By Proposition 16.14, we have
⊥
im(D − λ) = ker(D − λ)∗ = ker(D∗ − λ).
52 CHI-WAI LEUNG
Lemma 18.1. Let T ∈ B(H) be a normal operator (recall that T ∗ T = T T ∗ ). Then T is invertible
in B(H) if and only if there is c > 0 such that ∥T x∥ ≥ c∥x∥ for all x ∈ H.
Proof. The necessary part is obvious.
Now we want to show the converse. We first show the case where T is selfadjoint. Clearly, T
is injective from the assumption. By the Open Mapping Theorem, we need to show that T is
surjective.
⊥
In fact since ker T = imT ∗ and T = T ∗ , we see that the image of T is dense in H.
Now if y ∈ H, then there is a sequence (xn ) in H such that T xn → y. Thus, (T xn ) is a Cauchy
sequence. From this and the assumption give us that (xn ) is also a Cauchy sequence. If xn converges
to x ∈ H, then y = T x. Therefore the assertion is true when T is selfadjoint.
Now if T is normal, then we have ∥T ∗ x∥ = ∥T x∥ ≥ c∥x∥ for all x ∈ H by Proposition 16.11(ii).
Therefore, we have ∥T ∗ T x∥ ≥ c∥T x∥ ≥ c2 ∥x∥. Hence T ∗ T still satisfies the assumption. Note that
T ∗ T is selfadjoint. Therefore, we can apply the previous case to know that T ∗ T is invertible. This
implies that T is also invertible because T ∗ T = T T ∗ .
The proof is complete. □
S2
If we let R := I − Nc , then (I − R)−1 exists in B(H) and hence we have
∞
S 2 −1 X S2 n
( ) = (I − R)−1 = (I − ) .
Nc Nc
n=0
Then the result follows from
∞
1 X (T ∗ T )2 n ∗ ∗
T −1 = (I − ) T TT .
Nc Nc
n=0
□
Remark 18.6. In Proposition 18.5, we have shown that if T is selfadjoint, then σ(T ) ⊆ R. How-
ever, the converse does not hold. For example, consider H = C2 and
0 1
T = .
0 0
Example 18.7. Notice that the multiplication defines an isometry M : x ∈ ℓ∞ 7→ M (x) ∈ B(ℓ2 )
by M (x)(ξ)(n) := x(n)ξ(n); n = 1, 2... for ξ ∈ ℓ2 . Then M (x̄) = M (x)∗ for x ∈ ℓ∞ , and so, M (x)
is self-adjoint if and only if x is a R-sequence. Now let x ∈ ℓ∞ be a R-sequence. For simply for
each element x ∈ ℓ∞ , we also write x for M (x) as an element in B(ℓ2 ).
Now we claim that if x ∈ ℓ∞ is self-adjoint, then λ ∈ σ(x) if and only if inf |x(n) − λ| = 0.
n
Consequently, σp (x) = {xn : n = 1, 2...} and σ(x) = {x(n) : n = 1, 2....}.
In fact, for showing (⇐), let λ ∈ R such that inf |x(n) − λ| = 0. If x − λ is invertible in B(ℓ2 ), then
n
by Lemma 18.1, there is c > 0 such that ∥(x − λ)ξ∥ ≥ c for all ξ ∈ ℓ2 of norm one. In particular,
for each n = 1, 2..., we have |x(n) − λ| = ∥(x − λ)(en )∥ ≥ c > 0. It leads to a contradiction.
For showing (⇒), let λ ∈ R such that c := inf |x(n) − λ| > 0. Then x(n) ̸= λ for all n = 1, 2...
n
This implies that x − λ is injective. On the other hand, for any η ∈ ℓ2 , if (x(n) − λ)ξ(n) = η(n)
η(n)
for all n, then we have ξ(n) = x(n)−λ and so, |ξ(n)| ≤ |η(n)|
c . This gives ξ ∈ ℓ2 . Therefore, x − λ is
surjective and thus, x − λ is invertible. Hence, λ ∈/ σ(x).
From this, the last assertion follows because λ ∈ σ(x) if and only if λ = xn for some n or there is
a subsequence (xnk ) of (xn ) that converges to λ.
54 CHI-WAI LEUNG
Theorem 19.1. Liouville’s Theorem Every bounded entire function is a constant function.
Theorem 19.2. Using the notion as before, let T ∈ B(X). Then the spectrum σ(T ) ̸= ∅.
Proof. Assume that σ(T ) = ∅. Fix f ∈ B(X)∗ , define the map g(z) := f ((z − T )−1 ) is defined
for all z ∈ C. Note that g is continuous on C by considering the composition λ ∈ C 7→ λ − T 7→
55
(λ − T )−1 ∈ B(X) and using Proposition 17.5 (iv). Moreover, we have limz→∞ |g(z)| = 0. Thus, g
is a bounded function on C. On the other hand, if we fix a point c ∈ C, then we see that
g(z) − g(c)
lim = −f ((c − T )−2 ).
z→c z−c
Therefore, g is a bounded entire function. By the Liouville’s Theorem, f ((z − T )−1 ) is a constant
function on C. Then the Hahn-Banach Theorem implies that the function z ∈ C 7→ (z − T )−1 ∈
B(X) is constant on C. It leads to a contradiction. □
Proposition 20.1. Let S, T ∈ B(H) such that ST = T S. If S, T both are positive operators, then
so is ST .
S
Proof. If S = 0, then the assertion is clear. Now we assume that S ̸= 0. Put S1 := ∥S∥ . Set
Sn+1 := Sn − Sn2
for n = 1, 2, ....
Claim 1: 0 ≤ Sn ≤ I for all n = 1, .... The assertion will be obtained by induction on n.
Notice that as n = 1, clearly we have 0 ≤ S1 ≤ I. Suppose that the Claim 1 is true for n,
i.e., 0 ≤ Sn ≤ I and thus, we have 0 ≤ I − Sn ≤ I. This implies that for all x ∈ H we
have (Sn2 (I − Sn )x, x) = ((I − Sn )Sn x, Sn x) ≥ 0. This gives Sn2 (I − Sn ) ≥ 0. Similarly, we have
Sn (I − Sn )2 ≥ 0. Hence, we have 0 ≤ Sn2 (I − Sn ) + Sn (I − Sn )2 = Sn − Sn2 = Sn+1 . On the other
hand, we have 0 ≤ (I − Sn ) + Sn2 = I − Sn+1 because Sn2 ≥ 0 and I − Sn ≥ 0. Therefore Claim 1
follows from the induction.
The proof will be complete if we show that (ST x, x) ≥ 0 for all x ∈ H.
In fact, notice that we have
S1 = S12 + S2 = S12 + S22 + S3 = · · · = S12 + · · · + Sn2 + Sn+1 .
This implies that
S12 + · · · + Sn2 = S1 − Sn+1 ≤ S1
for all n = 1, 2.. because Sn+1 ≥ 0. Thus, we have
Xn Xn n
X
2
∥Sk x∥ = (Sk x, Sk x) = (Sk2 x, x) ≤ (S1 x, x)
k=1 k=1 k=1
P∞
for all n. This gives k=1 ∥Sk x∥2 < ∞ and so, Sn x → 0. This implies that
n
X
( Sk2 )x = S1 (x) − Sn+1 (x) → 0
k=1
P∞ 2
for all x ∈ H and so we have k=1 Sk (x) = S1 (x) for all x ∈ H. Finally, we complete the proof by
the following
∞
X ∞
X
2
(ST x, x) = ∥S∥(T S1 x, x) = ∥S∥ (T Sk x, x) = ∥S∥ (T Sk x, Sk x) ≥ 0
k=1 k=1
for all x ∈ H. □
56 CHI-WAI LEUNG
Proposition 20.2. Let Tn , n = 1, 2, .... and K be the bounded linear selfadjoint operators on H.
Suppose that
(1) T1 ≤ T2 ≤ · · · ≤ K.
(2) Tn Tm = Tn Tm and KTn = Tn K for all m, n = 1, 2.....
Then there is a bounded selfadjoint operator T on H with T ≤ K such that lim Tn x = T x for all
x ∈ H.
Proof. Now let Sn := K − Tn for n = 1, 2, ... Then 0 ≤ Sn for all n = 1, 2, .... By using Proposition
2 − S S = (S − S )S ≥ 0 and hence, S 2 ≥ S S for n ≥ m. Similarly, we
20.1, we see that Sm n m m n m m n m
also have Sn Sm ≥ Sn2 for n ≥ m. Therefore, we have
2
(20.1) Sm ≥ Sn Sm ≥ Sn2
for n ≥ m. Thus, ((Sm 2 x, x))∞ 2
m=1 is a decreasing sequence of non-negative numbers and so lim(Sn x, x)
exists for all x ∈ H. Moreover since Sn and Sm commutes to each other, Eq 20.1 gives
∥Sm x − Sn x∥2 = ((Sm − Sn )2 x, x)
2 2
= (Sm x, x) − 2(Sm Sn x, x) + (Sm x, x)
2
≤ (Sm x, x) − (Sn2 x, x) → 0
for n ≥ m and for all x ∈ H. This implies that (Sn x) is a Cauchy sequence and hence, lim Sn x
exists for all x ∈ H. This implies that T (x) := lim Tn (x) = K − lim Sn x exists for all x ∈ H. The
Uniform Boundedness Theorem tells us that T ∈ B(H). In addition T is selfadjoint because each
Tn is selfadjoint. The proof is complete. □
Theorem 20.3. If T is a bounded positive operator on H, then there is a unique positive operator
S such that S 2 = T . In this case, we call S the square root of T .
Proof. We show the existence first.
T
Clearly, we may assume that T ̸= 0 and T ≤ I by considering the operator ∥T ∥ . Put S0 = 0 and
1 2
Sn = Sn−1 + (T − Sn−1 )
2
for n = 1, 2, ..... Then Sn is a polynomial of T and so, all Sn ’s are selfadjoint operators and commute
to each other. Notice that since 0 < T ≤ I and by the definition of Sn , we have
1 2 1 1
I − Sn = I − Sn−1 − (T − Sn−1 ) = (I − Sn−1 )2 + (I − T ) ≥ 0.
2 2 2
Thus Sn ≤ I for all n = 0, 1, 2.... On the other hand, we have
1 1 1
(20.2) Sn+1 − Sn = Sn + (T − Sn2 ) − Sn−1 − (T − Sn−1 2
) = (Sn − Sn−1 )(I − (Sn + Sn−1 ))
2 2 2
1
for all n = 0, 1, 2.... Since Sn ≤ I, I − 2 (Sn + Sn−1 ) ≥ 0. Using Proposition 20.1 and the Eq 20.2,
we can apply induction on n to see that 0 = S0 ≤ · · · ≤ Sn ≤ Sn+1 ≤ · · · ≤ I for all n = 0, 1, 2....
Proposition 20.2 tells us that Sx := lim Sn x exists for all x ∈ H and S ∈ B(H). In addition S is
positive because Sn ≥ 0 for all n = 0, 1, 2.... Also, since Sn x = Sn−1 x + 21 (T − Sn−1
2 )x for all x ∈ H,
2
by taking n → ∞, we see that T x = S x for all x. Thus the operator S is as desired.
Finally, we show the uniqueness.
Now let R be another positive bounded operator on H such that R2 = T . Notice that RT = R3 =
T R. This implies that RS = SR because S is the ∥ · ∥-limit of the polynomials of T by the above
construction of S. Now we take any x ∈ H and put y := (S − R)x. Then we have
0 ≤ (Sy, y) + (Ry, y) = ((S + R)(S − R)x, y) = ((S 2 − R2 )x, y) = 0.
57
This implies that (Sy, y) = (Ry, y) = 0 because both are non-negative numbers. On the other
hand, since S ≤ 0, by above there is another positive operator W such that W 2 = S, and so we
have 0 = (Sy, y) = (W y, W y) that gives Sy = 0. Similarly, we also have Ry = 0. Finally, we have
∥(S − R)x∥2 = ((S − R)2 x, x) = ((S − R)y, x) = 0.
Thus, S = R as desired. The proof is complete. □
Definition 21.1. A linear operator T : H → H is said to be compact if for every bounded sequence
(xn ) in H, (T (xn )) has a norm convergent subsequence.
Write K(H) for the set of all compact operators on H and K(H)sa for the set of all compact
selfadjoint operators.
Remark 21.2. Let U be the closed unit ball of H. Clearly, T is compact if and only if the norm
closure T (U ) is a compact subset of H. Thus if T is compact, then T is bounded automatically
because every compact set is bounded. In particular, if T has finite rank, that is dim imT < ∞,
then T must be compact because every closed and bounded subset of a finite dimensional normed
space is compact. In addition, clearly we have the following result.
Proposition 21.3. The identity operator I : H → H is compact if and only if dim H < ∞.
X X 4L
|ym (k) − y(k)|2 < <ε
k2
k≥N k≥N
for all m. On the other hand, since lim ym (k) = y(k) for all k, we can choose a positive integer
m→∞
M such that
N
X −1
|ym (k) − y(k)|2 < ε
k=1
for all m ≥ M . Finally, we have ∥ym − y∥22 < 2ε for all m ≥ M .
Theorem 21.5. Let T ∈ B(H). Then T is compact if and only if T maps every weakly convergent
sequence in H to a norm convergent sequence.
Proof. We first assume that T ∈ K(H). Let (xn ) be a weakly convergent sequence in H. Since H is
reflexive, (xn ) is bounded by the Uniform Boundedness Theorem. Thus we can find a subsequence
(xj ) of (xn ) such that (T xj ) is norm convergent. Let y := limj T xj . We claim that y = limn T xn .
Suppose that y ̸= limn T xn . Then by the compactness of T again, we can find a subsequence (xi )
of (xn ) such that T xi converges to y ′ with y ̸= y ′ . Thus there is z ∈ H such that (y, z) ̸= (y ′ , z).
58 CHI-WAI LEUNG
On the other hand, if we let x be the weakly limit of (xn ), then (xn , w) → (x, w) for all w ∈ H.
Thus we have
(y, z) = lim(T xj , z) = lim(xj , T ∗ (z)) = (x, T ∗ z) = (T x, z).
j j
Similarly, we also have (y ′ , z) = (T x, z) and hence (y, z) = (y ′ , z) that contradicts to the choice of
z.
For the converse, let (xn ) be a bounded sequence. Then by Theorem 15.12, (xn ) has a weakly
convergent subsequence. Thus T (xn ) has a norm convergent subsequence by the assumption. Thus
T is compact. □
Example 21.7. Let k(z, w) ∈ C(T × T). Define an operator T : L2 (T) → L2 (T) by
Z
T ξ(z) := k(z, w)ξ(w)dw
T
for z ∈ T and ξ ∈ L2 (T). Then T is a compact operator.
Proof. Clearly, we have ∥T ∥ ≤ ∥k∥∞ . On the other hand, Stone-Weiestrass Theorem tells us the
polynomials of (z, z̄; w, w̄) are ∥ · ∥∞ -dense in C(T × T). Therefore, by using Proposition 21.6, it
suffices to show for the case k(z, w) = N i j
P
i,j=1 aij (z, z̄)w w̄ where aij (z, z̄) is a polynomial of (z, z̄)
of degree N . From this, we have
XN Z
T ξ(z) = aij (z, z̄) wi w̄j ξ(w)dw
i,j=1 T
for ξ ∈ L2 (T). Thus, T (ξ) ∈ span{z i z̄ j : 0 ≤ i, j ≤ N } which is of finite dimension for all ξ ∈ L2 (T).
This implies that T has finite dimensional range and thus, T is compact. The proof is complete. □
Proposition 21.9. Let T ∈ K(H) and let c ∈ C with c ̸= 0. Then T − c has a closed range.
1 1
Proof. Note that T ∈ K(H). Thus if we consider T −I, we may assume that c = 1. Let S = T −I.
c c
Let (xn ) be a sequence in H such that Sxn → x ∈ H in norm. By considering the orthogonal
decomposition H = ker S ⊕ (ker S)⊥ , we write xn = yn ⊕ zn for yn ∈ ker S and zn ∈ (ker S)⊥ . We
first claim that (zn ) is bounded. Suppose that (zn ) is unbounded. By considering a subsequence
zn
of (zn ), we may assume that we may assume that ∥zn ∥ → ∞. Put vn := ∈ (ker S)⊥ .
∥zn ∥
Since Szn = Sxn → x, we have Svn → 0. On the other hand, since T is compact, and (vn ) is
bounded, by passing a subsequence of (vn ), we may also assume that T vn → w. Since S = T − I,
vn = T vn − Svn → w − 0 = w ∈ (ker S)⊥ . In addition from this we have Svn → Sw. On the other
hand, we have Sw = limn Svn = limn T vn − limn vn = w − w = 0. Thus w ∈ ker S ∩ (ker S)⊥ . It
follows that w = 0. However, since vn → w and ∥vn ∥ = 1 for all n. It leads to a contradiction.
Thus (zn ) is bounded.
Finally we are going to show that x ∈ imS. Now since (zn ) is bounded, (T zn ) has a convergent
subsequence (T znk ). Let limk T znk = z. Then we have
znk = Sznk − T znk = Sxnk − T znk → x − z.
It follows that x = limk Sxnk = limk Sznk = S(x − z) ∈ imS. The proof is complete. □
Remark 21.10. Proposition 21.9 dose not hold as c = 0. Example 21.4 gives a counter example
about this. In fact, we can directly check that the vector y := (1, 21 , 13 , .....) ∈ im(T ) \ im(T ), where
T is the compact operator defined in Example 21.4.
Theorem 21.11. Fredholm Alternative Theorem : Let T ∈ K(H)sa and let 0 ̸= λ ∈ C. Then
T − λ is injective if and only if T − λ is surjective.
Proof. Since T is selfadjoint, σ(T ) ⊆ R. Thus if λ ∈ C \ R, then T − λ is invertible. Thus the result
holds automatically.
Now consider the case λ ∈ R \ {0}.
Then T − λ is also selfadjoint. From this and Proposition 16.14, we have ker(T − λ) = (im(T − λ))⊥
and (ker(T − λ))⊥ = im(T − λ).
Thus the proof is complete immediately by using Proposition 21.9. □
Corollary 21.12. Let T ∈ K(H)sa . Then we have σ(T ) \ {0} = σp (T ) \ {0}. Consequently if
the values m(T ) and M (T ) which are defined in Theorem 18.8 are non-zero, then both are the
eigenvalues of T and ∥T ∥ = max |λ|.
λ∈σp (T )
Proof. It follows immediately from the Fredholm Alternative Theorem. This, together with Theo-
rem 18.8, implies the last assertion. □
Example 21.13. Let T ∈ B(ℓ2 ) be defined as in Example 21.4. We have shown that T ∈ K(ℓ2 ) and
T is selfadjoint. Then by Corollary 21.12 and Corollary 21.8, we see that σ(T ) = {0, 1, 12 , 13 , .....}.
Lemma 21.14. Let T ∈ K(H)sa and let Eλ := {x ∈ H : T x = λx} for λ ∈ σ(T ) \ {0}, that is the
eigenspace of T corresponding to λ. Then dim Eλ < ∞.
Proof. It is because the restriction T |Eλ : Eλ → Eλ is also a compact operator on Eλ , then
dim Eλ < ∞ for all λ ∈ σ(T ) \ {0} = σp (T ) \ {0}. □
60 CHI-WAI LEUNG
Theorem 21.15. Let T ∈ K(H)sa . Suppose that dim H = ∞. Then σ(T ) = {λk : k = 1, ..., N } ∪
{0}, where 1 ≤ N ≤ ∞ and (λn ) is a sequence of non-zero real numbers with |λ1 | ≥ |λ2 | ≥ · · · and
λi ̸= λj for i ̸= j. Moreover, if (λn ) is an infinite sequence, then |λn | ↓ 0.
Proof. Note that since dim H = ∞, 0 ∈ σ(T ). In addition we have ∥T ∥ = max(|M (T )|, |m(T )|)
and σ(T ) \ {0} = σp (T ) \ {0}. Thus by Corollary 21.12, there is |λ1 | = max |λ| = ∥T ∥. Since
λ∈σp (T )
dim Eλ1 < ∞, then Eλ⊥1 ̸= 0. By considering the restriction of ⊥
T2 := T |Eλ1 , if T2 ̸= 0, then there is
0 ̸= |λ2 | = maxλ∈σp (T2 ) |λ| = ∥T2 ∥. Note that λ2 ∈ σp (T ) and |λ2 | ≤ |λ1 | because ∥T2 ∥ ≤ ∥T ∥. To
repeat the same step, if TN +1 = 0 for some N , then 0 ∈ σp (T ). Otherwise, we can get an infinite
sequence (λn ) such that (|λn |) is decreasing.
Now we claim that if (λn ) is an infinite sequence, then limn |λn | = 0.
Otherwise, there is η > 0 such that |λn | ≥ η for all n. If we let vn ∈ Eλn with ∥vn ∥ = 1 for all
n. Note that since dim H = ∞ and dim Eλ < ∞, for any λ ∈ σp (T ) \ {0}, there are infinite many
λn ’s. Then wn := |λ1n | vn is a bounded sequence and ∥T wn − T wm ∥2 = ∥vn − vm ∥2 = 2 for m ̸= n.
This is a contradiction since T is compact. Thus limn |λn | = 0.
Finally we need to check σ(T ) = {λ1 , λ2 , ...} ∪ {0}.
In fact, let µ ∈ σp (T ). Since |λ1 | = ∥T ∥ ≥ |µ|, |λm+1 | < |µ| ≤ |λm |. Note that we have Eα ⊥Eβ
for α and β in σp (T ) with α ̸= β. Then by the construction of λn ’s, we have µ = λm . For
example, if |λ2 | < |µ| ≤ |λ1 | and µ ̸= λ1 , then Eµ ⊥Eλ1 . Hence, we have Eµ ⊆ (Eλ1 )⊥ . Then
by the construction of λ2 , that is |λ2 | = ∥T2 ∥ ≥ |µ| which leads to a contradiction. Thus, if
|λ2 | < |µ| ≤ |λ1 |, then µ = λ1 . The proof is complete. □
Theorem 21.16. Spectral Decomposition Theorem: Let T ∈ K(H)sa and let (λn )N n=1 , (1 ≤
N ≤ ∞), be a sequence of eigenvalues of T given as in Theorem 21.15. For each λ ∈ σp (T ) \ {0},
put d(λ) := dim Eλ < ∞. Let {eλ,i : i = 1, ..., d(λ)} be an orthonormal basis for Eλ . Then we have
the following orthogonal decomposition:
N
M
(21.1) H = ker T ⊕ Eλn
n=1
∞
M
In here as N = ∞, the space Eλn means that the closed linear spanned by the set {xk : xk ∈
n=1
Eλk }.
Moreover B := {eλ,i : λ ∈ σp (T ) \ {0}; i = 1, .., d(λ)} forms an orthonormal basis of T (H), and we
have
N d(λ
X X n)
the decomposition (21.1). Moreover, from this we see that the family B forms an orthonormal basis
of (ker T )⊥ . On the other hand, we have (ker T )⊥ = imT ∗ = imT . Therefore, B is an orthonormal
basis for T (H) and Equation 21.2 follows.
For the last assertion,Pit needs to show that the series ∞
P
n=1 λn Pn converges to T in norm. Note
that if we put Sm := m n=1 λ n P n , then by the decomposition (21.1), lim Sm x = T x for all x ∈ H.
m→∞
Thus it suffices to show that (Sm )∞
m=1 is a Cauchy sequence in B(H). In fact we have
Corollary 21.17. T ∈ K(H) if and only if T can be approximated by finite rank operators.
Proof. The sufficient condition follows immediately from Proposition 21.6.
Conversely, for a general compact operator T , we can consider the decomposition:
1 1
T = (T + T ∗ ) + i( (T − T ∗ )).
2 2i
Note that Re(T ) := 12 (T + T ∗ ) (call the real part of T ) and Im(T ) := 2i
1
(T − T ∗ ) (call the imaginary
part of T ) both are the self-adjoint compact operators. From this, we see that the T can be
approximated by finite ranks operators by using Theorem 21.16. □
Proposition 22.2. Let S, T be the operators on H. Assume that T , S and ST are densely defined.
Then T ∗ S ∗ ⊂ (ST )∗ .
62 CHI-WAI LEUNG
Proof. We first claim that T ∗ S ∗ ⊂ (ST )∗ . Let x ∈ D(ST ) and y ∈ D(T ∗ S ∗ ). Then S ∗ y is defined
and S ∗ y ∈ D(T ∗ ). Since x ∈ D(ST ) we have x ∈ D(T ) and T x ∈ D(S). Thus we have
(ST x, y) = (T x, S ∗ y) = (x, T ∗ S ∗ y).
This implies that y ∈ D(ST )∗ and (ST )∗ (y) = T ∗ S ∗ y and hence T ∗ S ∗ ⊂ (ST )∗ . □
Example 22.3. First we recall that a function f : [a, b] → C is called an indefinite integral if there
is an element φ ∈ L1 [a, b] such that
Z x
f (x) = f (a) + φ(t)dt
a
for all x ∈ [a, b], where dt is the Lebesgue measure on [a, b]. In this case we have f ′ (x) = φ(x)
almost everywhere in (a, b).
Let
D := {f : [a, b] → C : f is an indefinite integral with f (a) = f (b) and f ′ ∈ L2 [a, b]}.
Note that D is dense subspace of L2 [a, b]. Define an operator T with D(T ) = D by
T f := if ′ .
for f ∈ D. We claim that T is self-adjoint. The proof is divided by several steps.
Claim 1: T ⊂ T ∗ .
In fact, let f, g ∈ D. Then we have
Z b
(T f, g) = if ′ (t)g(t)dt
a
Z b
= ig(t)df (t)
a
(22.1) Z b
= if (t)g(t)|ba −i f (t)g ′ (t)dt
a
Z b
= f (t)ig ′ (t)dt = (f, T g).
a
for all f ∈ D. This implies that (g + iΦ)⊥im(T ). If we let 1 ∈ L2 [a, b] be the function of constant
one in [a, b], then we have
Z b
(T f, 1) = if ′ (t)dt = i(f (b) − f (a)) = 0
a
for all f ∈ D, hence C1⊥im(T ). On the other hand, note that for any ξ ∈ L2 [a, b] if we put
Rb Rb Rx
ξ1 = ξ − a ξ(t)dt ∈ L2 [a, b], then a ξ1 (t)dt = 0. Let h(x) := i a ξ1 (t)dt. Then h ∈ D and
T h = ξ1 . Therefore, we have L2 [a, b] = C1+im(T ) and hence we have the orthogonal decomposition
L2 [a, b] = C1 ⊕ im(T ). In particular, (im(T ))⊥ = C1. This implies that g + iΦ = c for some
constant c. Then g ′ = −iΦ′ = −iφ ∈ L2 [a, b], so g is an indefinite integral because g ′ ∈ L1 [a, b].
Moreover, we see that g(b) = g(a) = c because Φ(b) = Φ(a) = 0. We can now conclude that g ∈ D.
The proof is complete.
Proposition 22.6. Using the notation as above, let T be a densely operator on H. Then G(T ∗ ) =
(V (G(T )))⊥ . Consequently, the adjoint operator T ∗ is closed. In particular, if T is self-adjoint,
then T is closed.
Proof. Note that for x ∈ D(T ∗ ) and y ∈ D(T ), we have ((x, T ∗ x), V (y, T y)) = 0 Therefore, we
have G(T ∗ ) ⊆ (V (G(T )))⊥ . On the other hand, if (u, v)⊥(−T y, y) for all y ∈ D(T ). Then we
have (v, y) = (u, T y) and hence, u ∈ D(T ∗ ) and T ∗ u = v. Therefore, (u, v) ∈ G(T ∗ ). The proof is
complete. □
Proposition 22.7. Let T be a symmetric operator on H. Then the following statements are
equivalent.
(i) T is self-adjoint.
(ii) T is closed and ker(T ∗ ± i) = {0}.
(iii) im(T ± i) = H.
64 CHI-WAI LEUNG
Proof. For (i) ⇒ (ii), assume that T is self-adjoint. Then by Proposition 22.6, T is closed. Next
we show ker(T ∗ − i) = {0}. Let y ∈ D(T ∗ ) such that T ∗ y = iy. Since D(T ) = D(T ∗ ), we have
i(y, y) = (T y, y) = (y, T ∗ y) = −i(y, y), so y = 0. Similarly, we have ker(T ∗ + i) = {0}.
For (ii) ⇒ (iii), we first claim that im(T + i) is dense in H. Let z⊥im(T + i). Then z⊥(T + i)x for
all x ∈ D(T ), and thus we have (T x, z) = (x, −iz). This implies that z ∈ D(T ∗ ) and T ∗ z = −iz.
Thus, z ∈ ker(T ∗ + i), so z = 0. Therefore, it suffices to show that im(T + i) is closed. Let (xn ) be
a sequence in D(T ) such that lim(T − i)xn = y. Since T is symmetric, we have
∥T (xm − xn ) + i(xm − xn )∥2 = ∥T (xm − xn )∥2 + ∥(xm − xn )∥2
for all m, n. From this we see that u := lim xn and v := lim T xn both exist. T is closed by the
assumption, so u ∈ D(T ) and T u = v. Therefore, we have
y = lim(T xn + ixn ) = v + iu = (T + i)u ∈ im(T + i).
Hence im(T + i) = H. Similarly, we have im(T − i) = H.
For the last implication (iii) ⇒ (i), since T ⊂ T ∗ , we need to show that D(T ∗ ) ⊆ D(T ). Let
u ∈ D(T ∗ ). Since im(T − i) = H, there is an element v ∈ D(T ) such that
(T − i)v = (T ∗ − i)u.
Since T ⊂ T ∗ , we have (T − i)v = (T ∗ − i)v, thus, v − u ∈ ker(T ∗ − i). Then for any z ∈ D(T ), we
have
((T + i)z, v − u) = (z, (T + i)∗ (v − u)) = (z, (T ∗ − i)(v − u)) = 0.
im(T + i) = H by assumption, so u = v ∈ D(T ). The proof is complete. □
Proposition 22.8. Let T be a symmetric operator on H. Then there is the smallest closed exten-
sion of T , denoted it by T . We call T the closure of T . In addition, G(T ) = G(T ) and T = T ∗∗ .
Proof. Let D(T ) := {x ∈ H : (x, y) ∈ G(T ) for some y ∈ H}. We first note for each element
x ∈ D(T ), there is a unique element y ∈ H so that (x, y) ∈ G(T ). In fact, if (x, y) ∈ G(T ), there is
a sequence (xn ) in D(T ) such that lim xn = x and lim T xn = y. Note that for any u ∈ D(T ), since
T is symmetric, we have
(T u, x) = lim(T u, xn ) = lim(u, T xn ) = (u, y).
Therefore, y is uniquely determined by x because D(T ) is dense in H. Hence, we can define T x = y
for x ∈ D(T ). Clearly, we have G(T ) = G(T ) by the construction of T , and hence T is closed.
Moreover, we can directly show that T is the smallest closed extension of T .
For the last assertion, since T ⊂ T ∗ , T ∗ is densely defined, so T ∗∗ := (T ∗ )∗ is defined. Since
V 2 = −I and V is an isometry and an orthogonal preserver, by using Proposition 22.6, we have
G(T ∗∗ ) = [V G(T ∗ )]⊥
= V [G(T ∗ )⊥ ]
= V [V (G(T ))]
= V 2 (G(T ))
= G(T ).
Thus, T = T ∗∗ . □
References
[1] L.V. Ahlfors, Complex analysis, 3rd edition, Mc Graw Hill, (1979).
[2] S. Banach, Theory of linear operators, English translation by F. Jallett, North-Holland, (1987).
[3] A. Bowers and N.J. Kalton, An introductory course in functional analysis, Springer, (2014).
[4] N.L. Carothers, A short course on Banach space theory, Cambridge University Press (2005).
65
[5] W. Cheney, Ward and A. Goldstein, Proximity maps for convex sets. Proc. Amer. Math. Soc. 10 (1959), 448–450.
[6] J. Diestel, Sequences and series in Banach spaces, Springer-Verlag, (1984).
[7] P.R. Halmos, Naive set theory, Dover publications, (2017).
[8] R.C. James, A non-refexivity of Banach space isometric with its second conjugate space, Proc. N.A.S. USA, 37
(1951), 174-177.
[9] Y. Katznelson, An introduction to Harmonic analysis, 3rd edition, Cambridge Univ Press, (2004).
[10] R.E. Megginson, An introduction to Banach space theory, GTM vol. 183, Springer, (1998).
(Chi-Wai Leung) Department of Mathematics, The Chinese University of Hong Kong, Shatin, Hong
Kong
Email address: cwleung@math.cuhk.edu.hk