Functional Analysis
Functional Analysis
CHI-WAI LEUNG
Throughout this note, all spaces X, Y, .. are normed spaces over the field K = R or C. Let
BX := {x ∈ X : kxk ≤ 1} and SX := {x ∈ X : kxk = 1} denote the closed unit ball and the unit
sphere of X respectively.
Example 1.4. 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, that is, 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, that is,
kf k∞ = 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 2.2. All norms on a finite dimensional vector space are equivalent.
Lemma 2.4. 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{kx0 − yk : y ∈ Y } ≥ θ.
3
Remark 2.5. 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 [2, 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. Notice 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 kx − yk∞ < 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 kxk∞ = 1, we can find 0 < a < 1/4 such that |x(t)| ≤ 1/4
for all t ∈ [0, a]. Notice that since x is uniform continuous on [a, 1], for any 0 < ε < 1/4, there
is δ > 0 such that |x(t) − x(t0 )| < ε/4 when |t − t0 | < δ. Now we find a partition a = t0 < t1 <
· · · < tn = 1 with tk − tk−1 < δ for all k = 1, 2, ...n and |x(tk )| < 1 for all k = 1, 2..., n − 1. Then
sup{|x(t) − x(t0 )| : t, t0 ∈ [tk−1 , tk ]} < ε/4. We let pk−1 := sup{t ∈ [tk−1 , tk ] : x|[tk−1 ,t] > −1 + ε} if
it exists, otherwise, put pk−1 := tk−1 . Similarly, let qk := inf{t ∈ [tk−1 , tk ] : x|[t,tk ] > −1 + ε} if it
exists, otherwise, put qk := tk . So, one can find a continuous function φ on [a, 1] such that
ε if t ∈ [tk−1 , tk ] and x|[tk−1 ,tk ] > −1 + ε.
−ε
if t ∈ [pk−1 , qk ] and x|[tk−1 ,tk ] ≯ −1 + ε.
φ(t) = −2ε
pk−1 −tk−1 (t − tk−1 ) + ε if x|[tk−1 ,tk ] ≯ −1 + ε and tk−1 < t < pk−1 .
2ε (t − t ) + ε
if x| ≯ −1 + ε and q < t < t .
tk −qk k [tk−1 ,tk ] k k
Notice that if x|[tk−1 ,tk ] ≯ −1 + ε, then tk−1 < pk−1 or qk < tk . So, kx|[a,1] − φk∞ < 1.
R1
It is because kφk∞ < 2ε, we have | a φ(t)dt| ≤ 2ε(1 − a). On the other hand, as |x(t)| < 1/4
on [0, a], so if we further 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) and
Ra R1
0 y1 (t)dt = − a φ(t)dt. Now we define y = y1 on [0, a] and y = φ on [a, 1]. Then ky − xk∞ < 1
and y ∈ Y is as desired.
Theorem 2.6. 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 2.3(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
kx1 − x2 k ≥ 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 kx3 − xk k ≥ 1/2 for
k = 1, 2. To repeat the same step, there is a sequence (xn ) ∈ SX such that kxm − xn k ≥ 1/2 for
all n 6= m. Thus, (xn ) is a bounded sequence without any convergence subsequence. So, BX is not
compact. The proof is finished.
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 2.6 implies the following corollary immediately.
Corollary 2.7. Let X be a normed space. Then X is locally compact if and only if dim X < ∞.
4 CHI-WAI LEUNG
Proposition 3.3. If X is of finite dimension normed space, then for any linear operator T from
X into a normed space Y must be bounded.
Proof. Let k · k0 be the equivalent norm on X defined as in the proof of Proposition 2.2. It is clear
that T is continuous at 0 with respect to the norm k · k0 . So, T is bounded by Proposition 3.1 at
once.
Proposition 3.4. 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
(3.1) kx̄k = inf{kx + yk : y ∈ Y }.
If we let π : X → X/Y be the natural projection, that is π(x) = x̄ for all x ∈ X, then (X/Y, k · k)
is a normed space and π is bounded with kπk ≤ 1. In particular, kπk = 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 k · k in (3.1) the quotient norm on X/Y .
Proof. Notice that since Y is closed, one can directly check that kx̄k = 0 if and only is x ∈ Y , that
is, x̄ = 0̄ ∈ X/Y . It is easy to check the other conditions of the definition of a norm. So, X/Y is
a normed space. Also, it is clear that π is bounded with kπk ≤ 1 by the definition of the quotient
norm on X/Y .
5
Furthermore, if Y ( X, then by using the Riesz’s Lemma 2.4, we see that kπk = 1 at once.
We are going to 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 (Why?).
Indeed, since (x̄n ) is a Cauchy sequence, we can find a subsequence (x̄nk ) of (x̄n ) such that
kx̄nk+1 − x̄nk k < 1/2k
for all k = 1, 2.... Then by the definition of quotient norm, there is an element y1 ∈ Y such that
kxn2 − xn1 + y1 k < 1/2. Notice that we have, xn1 − y1 = x̄n1 in X/Y . So, there is y2 ∈ Y such that
kxn2 −y2 −(xn1 −y1 )k < 1/2 by the definition of quotient norm again. Also, we have xn2 − y2 = x̄n2 .
Then we also have an element y3 ∈ Y such that kxn3 − y3 − (xn2 − y2 )k < 1/22 . To repeat the same
step, we can obtain a sequence (yk ) in Y such that
kxnk+1 − yk+1 − (xnk − yk )k < 1/2k
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, notice 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. So, (x̄nk ) is a convergent subsequence of (x̄n ) in X/Y . The proof is
complete.
Corollary 3.5. 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}, the kernel of T , is closed.
Proof. The necessary part is clear.
Now assume that ker T is closed. Then by Proposition 3.4, X/ ker T becomes a normed space.
Also, it is known that there is a linear injection Te : X/ ker T → Y such that T = Te ◦ π, where
π : X → X/ ker T is the natural projection. Since dim Y < ∞ and Te is injective, dim X/ ker T < ∞.
This implies that Te is bounded by Proposition 3.3. Hence T is bounded because T = Te ◦ π and π
is bounded.
Remark 3.6. The converse P of Corollary 3.5 does not hold when Y is of infinite dimension. For
example, let X := {x ∈ `2 : ∞ 2 2
n=1 n |x(n)| < ∞} (notice that X is a vector space Why?) and
2
Y = ` . Both X and Y are endowed with k · k2 -norm.
Define T : X → Y by T x(n) = nx(n) for x ∈ X and n = 1, 2.... Then T is an unbounded
operator(Check !!). Notice that ker T = {0} and hence, ker T is closed. So, the closeness of ker T
does not imply the boundedness of T in general.
We say that 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 also write
X = Y if X and Y are isometric isomorphic.
Recall that a metric space is said to be separable if there is a countable dense subset, for example,
the base field K is separable. Also, it is easy to see that the separability is preserved under a
homeomorphism.
Definition 3.7. We say that a sequence of element (en )∞ n=1 in a normed space X is called a
Schauder base 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.
6 CHI-WAI LEUNG
Remark 3.8. Notice that if X has a Scahuder base, then X must be separable. The following
natural question we first raised by Banach (1932).
The base problem: Does every separable Banach space have a Schauder base?
The answer is “No00 !
This problem was completely solved by P. Enflo in 1973.
Proposition 3.10. 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
kT k = sup{kT xk : x ∈ BX }.
be defined as in Proposition 3.1.
Then (B(X, Y ), k · k) becomes a normed space.
Furthermore, if Y is a Banach space, then so is B(X, Y ).
Proof. One can directly check that B(X, Y ) is a normed space (Do It By Yourself !).
We are going to show that B(X, Y ) is complete if Y is a Banach space. Let (Tn ) be a Cauchy
sequence in L(X, Y ). Then for each x ∈ X, it is easy to see that (Tn x) is also a Cauchy sequence
in Y . So, lim Tn x exists in Y for each x ∈ X because Y is complete. Hence, one can define a map
T x := lim Tn x ∈ Y for each x ∈ X. It is clear that T is a linear map from X into Y .
It needs to show that T ∈ L(X, Y ) and kT − Tn k → 0 as n → ∞. Let ε > 0. Since (Tn ) is a Cauchy
sequence in L(X, Y ), there is a positive integer N such that kTm − Tn k < ε for all m, n ≥ N . So, we
have k(Tm − Tn )(x)k < ε for all x ∈ BX and m, n ≥ N . Taking m → ∞, we have kT x − Tn xk ≤ ε
for all n ≥ N and x ∈ BX . Therefore, we have kT − Tn k ≤ 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 kT − Tn k → 0 as n → ∞.
Therefore, limn Tn = T exists in B(X, Y ).
4. Dual Spaces
By Proposition 3.10, we have the following assertion at once.
Proposition 4.1. Let X be a normed space. Put X ∗ = B(X, K). Then X ∗ is a Banach space and
is called the dual space of X.
Example 4.2. Let X = KN . Consider the usual Euclidean norm on X, that is, k(x1 , ..., xN )k :=
|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 . Notice that θx(y) = hx, yi, 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 )∗ .
7
for x ∈ `1 and η ∈ c0 .
Then T is isometric isomorphism and hence, c∗0 = `1 .
Example 4.4. We have the other important examples of the dual spaces.
(i) (`1 )∗ = `∞ .
(ii) For 1 < p < ∞, (`p )∗ = `q , 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 4.3 (see also in [3, Chapter
8]). Part (iii) is known as the Riesz representation Theorem which is referred to [3, Section 21.5]
for the details.
8 CHI-WAI LEUNG
5. Hahn-Banach Theorem
All spaces X, Y, Z... are normed spaces over the field K throughout this section.
Lemma 5.1. Let Y be a subspace of X and v ∈ X \ Y . Let Z = Y ⊕ Kv be the linear span of Y
and v in X. If f ∈ Y ∗ , then there is an extension F ∈ Z ∗ of f such that kF k = kf k.
Proof. We may assume that kf k = 1 by considering the normalization f /kf k if f 6= 0.
Case K = R:
We first note that since kf k = 1, we have |f (x) − f (y)| ≤ k(x + v) − (y + v)k for all x, y ∈ Y . This
implies that −f (x) − kx + vk ≤ −f (y) + ky + vk for all x, y ∈ Y . Now let γ = sup{−f (x) − kx + vk :
x ∈ X}. This implies that γ exists and
(5.1) −f (y) − ky + vk ≤ γ ≤ −f (y) + ky + vk
for all y ∈ Y . We define F : Z −→ R by F (y + αv) := f (y) + αγ. It is clear that F |Y = f . For
showing F ∈ Z ∗ with kF k = 1, since F |Y = f on Y and kf k = 1, it needs to show |F (y + αv)| ≤
ky + αvk for all y ∈ Y and α ∈ R.
In fact, for y ∈ Y and α > 0, then by inequality 5.1, we have
(5.2) |F (y + αv)| = |f (y) + αγ| ≤ ky + αvk.
Since y and α are arbitrary in inequality 5.2, we see that |F (y + αv)| ≤ ky + αvk for all y ∈ Y and
α ∈ R. Therefore the result holds when K = R.
Now for the complex case, let h = Ref and g = Imf . Then f = h + ig and f, g both are real linear
with khk ≤ 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 Z := Y ⊕ Rv ⊕ iRv of h such that kHk = khk. Now define F : Z −→ C by F (·) := H(·) − iH(i·).
Then F ∈ Z ∗ and F |Y = f . Thus it remains to show that kF k = kf k = 1. It needs to show
that |F (z)| ≤ kzk for all z ∈ Z. Note for z ∈ Z, 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)| ≤ kHkke−iθ zk ≤ kzk.
The proof is finished.
Remark 5.2. 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.
The following is the typical argument of applying the Zorn’s Lemma.
Theorem 5.3. Hahn-Banach Theorem : 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 kF k = kf k.
Proof. Let X be the collection of the pairs (Y1 , f1 ), where Y ⊆ Y1 is a subspace of X and f1 ∈ Y1∗
such that f1 |Y = f and kf1 kY1∗ = kf kY ∗ . 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 5.1 will give Ye = X. The proof is finished.
9
Proposition 5.4. Let X be a normed space and x0 ∈ X. Then there is f ∈ X ∗ with kf k = 1 such
that f (x0 ) = kx0 k. Consequently, we have
kx0 k = sup{|g(x)| : g ∈ BX ∗ }.
Also, if x, y ∈ X with x 6= y, then there exists f ∈ X ∗ such that f (x) 6= f (y).
Proof. Let Y = Kx0 . Define f0 : Y → K by f0 (αx0 ) := αkx0 k for α ∈ K. Then f0 ∈ Y ∗ with
kf0 k = kx0 k. So, the result follows from the Hahn-Banach Theorem at once.
Remark 5.5. Proposition 5.4 tells us that the dual space X ∗ of X must be non-zero. Indeed, the
dual space X ∗ is very “Large00 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 6= x2 , we can
find an element T ∈ B(X, Y ) such that T x1 6= T x2 . In fact, fix a non-zero element y ∈ Y . Then
by Proposition 5.4, there is f ∈ X ∗ such that f (x1 ) 6= f (x2 ). So, if we define T x = f (x)y, then
T ∈ B(X, Y ) as desired.
Proposition 5.6. With the notation as above, if M is closed subspace and v ∈ X \ M , then there
is f ∈ X ∗ such that f (M ) ≡ 0 and f (v) 6= 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. Notice that v̄ := π(v) 6= 0 ∈ X/M because v̄ ∈ X \ M . Then by
Corollary 5.4, there is a non-zero element f¯ ∈ (X/M )∗ such that f¯(v̄) 6= 0. So, the linear functional
f := f¯ ◦ π ∈ X ∗ is as desired.
Remark 5.8. The converse of Proposition 5.7 does not hold. For example, consider X = `1 . Then
`1 is separable but the dual space (`1 )∗ = `∞ is not.
Proposition 5.9. 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 kT ∗ k = kT k. In this case, T ∗ is called the
adjoint operator of T .
10 CHI-WAI LEUNG
Example 5.10. 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 4.2. 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 ∗ ∗ ∗ 0
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 = a0lk . Hence the
transpose (akl )t is the the representative matrix of T ∗ .
Proposition 5.11. 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 notice that for any x∗ ∈ X ∗ , the image i∗ x∗ in Y ∗ is just the restriction
of x∗ on Y , write x∗ |Y . Now let φ ∈ Y ∗∗ . Then for any x∗ ∈ X ∗ , we have
|i∗∗ φ(x∗ )| = |φ(i∗ x∗ )| = |φ(x∗ |Y )| ≤ kφkkx∗ |Y kY ∗ ≤ kφkkx∗ kX ∗ .
So, ki∗∗ φk ≤ kφk. It remains to show the inverse inequality. Now for each y ∗ ∈ Y ∗ , the Hahn-
Banach Theorem gives an element x∗ ∈ X ∗ such that kx∗ kX ∗ = ky ∗ kY ∗ and x∗ |Y = y ∗ and hence,
i∗ x∗ = y ∗ . Then we have
|φ(y ∗ )| = |φ(x∗ |Y )| = |φ(i∗ x∗ )| = |(i∗∗ ◦ φ)(x∗ )| ≤ ki∗∗ φkkx∗ kX ∗ = ki∗∗ φkky ∗ kY ∗
for all y ∗ ∈ Y ∗ . Therefore, we have ki∗∗ φk = kφk.
For Part (ii), let ψ ∈ (X/Y )∗ . Notice that since kπ ∗ k = kπk ≤ 1, we have kπ ∗ ψk ≤ kψk. On the
other hand, for each x̄ := π(x) ∈ X/Y with kx̄k < 1, we can choose an element m ∈ Y such that
kx + mk < 1. So, we have
|ψ(x̄)| = |ψ ◦ π(x)| = |ψ ◦ π(x + m)k ≤ kψ ◦ πk = kπ ∗ (ψ)k.
Thus we have kψk ≤ kπ ∗ (ψ)k. The proof is finished.
Remark 5.12. By using Proposition 5.11, we can give an alternative proof of the Riesz’s Lemma
2.4.
With the notation as in Proposition 5.11, if Y ( X, then we have kπk = kπ ∗ k = 1 because π ∗ is an
isometry by Proposition 5.11(ii). Thus we have kπk = sup{kπ(x)k : x ∈ X, kxk = 1} = 1. So, for
any 0 < θ < 1, we can find element z ∈ X with kzk = 1 such that θ < kπ(z)k = inf{kz+yk : y ∈ Y }.
The Riesz’s Lemma follows.
6. Reflexive Spaces
Proposition 6.1. 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.
11
Proof. Note that for x ∈ X and x∗ ∈ BX ∗ , we have |Q(x)(x∗ )| = |x∗ (x)| ≤ kxk. Then kQ(x)k ≤
kxk.
It remains to show that kxk ≤ kQ(x)k for all x ∈ X. In fact, for x ∈ X, there is x∗ ∈ X ∗ with
kx∗ k = 1 such that kxk = |x∗ (x)| = |Q(x)(x∗ )| by Proposition 5.4. Thus we have kxk ≤ kQ(x)k.
The proof is finished.
Example 6.6. By using Proposition 6.5, we immediately see that the space `∞ is not reflexive
because it contains a non-reflexive closed subspace c0 .
Proposition 6.7. Let X be a normed 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.
12 CHI-WAI LEUNG
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. So, 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. Notice that the adjoint operator π ∗ : (X/M )∗ → X ∗ is an isometry (Check
!). So, (X/M )∗ can be viewed as a closed subspace of X ∗ . So, by Part (i) and Proposition 6.5, we
see that (X/M )∗ is reflexive. Then X/M is reflexive by using Part (i) again.
The proof is complete.
Lemma 6.8. 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. It needs to show that
re is an isometry. Notice that re(x∗ + M ⊥ ) = x∗ |M for all x∗ ∈ X ∗ . Now for any x∗ ∈ X ∗ , we have
kx∗ + y ∗ kX ∗ ≥ kx∗ + y ∗ kM ∗ = kx∗ |M kM ∗ for all y ∗ ∈ M ⊥ . So we have ke
r(x∗ + M ⊥ )k = kx∗ |M kM ∗ ≤
∗ ⊥
kx + M k. It remains 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 kz ∗ k = kx∗ |M kM ∗ . Then x∗ − z ∗ ∈ M ⊥ and hence, we have x∗ + M ⊥ = z ∗ + M ⊥ .
This implies that
kx∗ + M ⊥ k = kz ∗ + M ⊥ k ≤ kz ∗ k = kx∗ |M kM ∗ = ke
r(x∗ + M ⊥ )k.
The proof is complete.
Proposition 6.9. (Three space property): Let M be a closed subspace of a normed space X.
If M and the quotient space X/M both are reflexive, then so is X.
Proof. Let π : X → X/M be the natural projection. Let ψ ∈ X ∗∗ . We going to show that
ψ ∈ im(QX ). Since π ∗∗ (ψ) ∈ (X/M )∗∗ , there exists x0 ∈ X such that π ∗∗ (φ) = QX/M (x0 + M )
because X/M is reflexive. So we have
π ∗∗ (ψ)(x̄∗ ) = QX/M (x0 + M )(x̄∗ )
for all x̄∗ ∈ (X/M )∗ . This implies that
ψ(x̄∗ ◦ π) = ψ(π ∗ x̄∗ ) = π ∗∗ (ψ)(x̄∗ ) = QX/M (x0 + M )(x̄∗ ) = x̄∗ (x0 + M ) = QX x0 (x̄∗ ◦ π)
for all x̄∗ ∈ (X/M )∗ . Therefore, we have
ψ = QX x 0 on M ⊥ .
So, we have ψ − QX (x0 ) ∈ X ∗ /M ⊥ . Let f : M ∗ → X ∗ /M ⊥ be the inverse of the isometric
isomorphism re which is defined as in Lemma 6.8. Then the composite (ψ − QX x0 ) ◦ f : M ∗ →
X ∗ /M ⊥ → K lies in M ∗∗ . Then by the reflexivity of M , there is an element m0 ∈ M such that
(ψ − QX x0 ) ◦ f = QM (m0 ) ∈ M ∗∗ .
On the other hand, notice that for each x∗ ∈ X ∗ , we can find an element m∗ ∈ M ∗ such that
f (m∗ )x∗ + M |bot ∈ X ∗ /M ⊥ because f is surjective, moreover, by the construction of re in Lemma
6.8, we see that x∗ |M = m∗ . This gives
ψ(x∗ ) − x∗ (x0 ) = (ψ − QX x0 )(m∗ ) ◦ f = QM (m0 )(m∗ ) = m∗ (m0 ) = x∗ (m0 ).
Thus, we have ψ(x∗ ) = x∗ (x0 +m0 ) for all x∗ ∈ X ∗ . From this we have ψ = QX (x0 +m0 ) ∈ im(QX )
as desired. The proof is complete.
13
Proposition 7.2. A weak limit of a sequence is unique if it exists. In this case, if (xn ) weakly
w
converges to x, write x = w-lim xn or xn −
→ x.
n
Proof. The uniqueness follows from the Hahn-Banach Theorem immediately.
w
Remark 7.3. It is clear that 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 = `1 but (en ) is not
convergent in c0 .
Proposition 7.4. 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 base for X and let fk be
the k-th coordinate functional corresponding to the base B, that is 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 . So, we have kxn − xk → 0.
Definition 7.5. 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 7.6. 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.
Example 7.7. Remark 7.3 has shown that the w-convergence does not imply k · k-convergence.
We now claim that the w∗ -convergence also Does Not imply the w-convergence.
Consider X = c0 . Then c∗0 = `1 and c∗∗ 1 ∗ ∞ ∗ 1 ∗
0 = (` ) = ` . Let en = (0, ...0, 1, 0...) ∈ ` = c0 , where
w∗
the n-th coordinate is 1. Then e∗n −−→ 0 but e∗n 9 0 weakly because e∗∗ (e∗n ) ≡ 1 for all n, where
e∗∗ := (1, 1, ...) ∈ `∞ = c∗∗ ∗
0 . Hence the w -convergence does not imply the w-convergence.
w
Proposition 7.8. 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:
k·k
(i) : fn −−→ f ;
w
(ii) : fn −→ f;
w∗
(iii) : fn −−→ f .
Theorem 7.9. (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
14 CHI-WAI LEUNG
convergent subsequence, say (f2,k (x2 )), of (f1,k (x2 )). Put c2 := limk f2,k (x2 ). Notice that we still
have c1 = limk f2,k (x1 ). To repeat the same step, if we define (m, k) ≤ (m0 , k 0 ) if m < m0 ; or
m = m0 with k ≤ k 0 , 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 ). Notice that for each i, we have limk hk (xi ) =
limk fi,k (xi ) = ci by the construction (ii) above. Since (khk k) 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
finished.
Remark 7.10. Theorem 7.9 does not hold if the separability of X is removed.
For example, consider X = `∞ and δn the n-th coordinate functional on `∞ . Then δn ∈ (`∞ )∗
with kδn k(`∞ )∗ = 1 for all n. Suppose that (δn ) has a w∗ -convergent subsequence (δnk )∞
k=1 . Define
∞
x ∈ ` by
0
if m 6= 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. So (δn ) has no
w∗ -convergent subsequence.
Corollary 7.11. Let X be a separable space. In X ∗ assume that the set of all w∗ -convergent
sequences coincides with the set of all normed convergent sequences, that is a sequence (fn ) is
w∗ -convergent if and only if it is norm convergent. Then dim X < ∞.
Proof. It needs to show that the closed unit ball BX ∗ in X ∗ is compact in norm. Let (fn ) be a
sequence in BX ∗ . By using Theorem 7.9, (fn ) has a w∗ -convergent subsequence (fnk ). Then by the
assumption, (fnk ) is norm convergent. Note that if lim fnk = f in norm, then f ∈ BX ∗ . So BX ∗ is
k
compact and thus dim X ∗ < ∞. So dim X ∗∗ < ∞ that gives dim X is finite because X ⊆ X ∗∗ .
Corollary 7.12. 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 notice that since X is reflexive and separable,
X ∗ is also separable by Proposition 5.7. So, we can apply Theorem 7.9, (Qxn ) has a w∗ -convergent
subsequence (QxnK ) in X ∗∗ = Q(X) and hence, (xnk ) is weakly convergent in X.
Lemma 8.2. 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.
15
Corollary 8.3. 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 8.2, 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 kx + mk < 1. Hence we have x + m ∈ U and
x̄ = π(x + m) ∈ π(U ).
Lemma 8.4. 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, that is,
U := {x ∈ X : kxk < 1}.
Proof. Set U (r) := {x ∈ X : kxk < 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 second category theorem, there
exists N such that int T (U (N )) 6= ∅. Let y 0 be an interior point of T (U (N )). Then there is
η > 0 such that BY (y 0 , η) ⊆ T (U (N )). Since BY (y 0 , η) ∩ T (U (N )) 6= ∅, we may assume that
y 0 ∈ T (U (N )). Let x0 ∈ U (N ) such that T (x0 ) = y 0 . Then we have
0 ∈ BY (y 0 , η) − y 0 ⊆ T (U (N )) − T (x0 ) ⊆ T (U (2N )) = 2N T (U (1)).
So we have 0 ∈ 2N1
(BY (y 0 , η) − y 0 ) ⊆ T (U (1)). Hence 0 is an interior point of T (U (1)). So Claim 1
follows.
Therefore there is r > 0 such that BY (0, r) ⊆ T (U (1)). This implies that we have
(8.1) 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 8.1, there is x1 ∈ U (1) such that ky − T (x1 )k < r/2. Then by using Eq 8.1
again, there is x2 ∈ U (1/2) such that ky − T (x1 ) − T (x2 )k < r/22 . To repeat the same steps, there
exists is a sequence (xk ) such that xk ∈ U (1/2k−1 ) and
ky − T (x1 ) − T (x2 ) − ... − T (xk )k < r/2k
for all k. On the other hand, since ∞
P P∞ k−1 and X is Banach, x :=
P∞
k=1 kxk k ≤ k=1 1/2 k=1 xk
exists in X and kxk ≤ 2. This implies that y = T (x) and kxk < 3.
Thus we the result follows.
Theorem 8.5. Open Mapping Theorem : Retains the notation as in Lemma 8.4. Then T is
an open mapping.
Proof. The proof is finished by using Lemmas 8.2 and 8.4 at once.
Proposition 8.6. Let T be a bounded linear isomorphism between Banach spaces X and Y . Then
T −1 must be bounded.
Consequently, if k · k and k · k0 both are complete norms on X such that k · k ≤ ck · k0 for some c > 0,
then these two norms k · k and k · k0 are equivalent.
16 CHI-WAI LEUNG
Proof. The first assertion follows from the Open Mapping Theorem at once.
Therefore, the last assertion can be obtained by considering the identity map I : (X, k·k) → (X, k·k0 )
which is bounded by the assumption.
Corollary 8.7. Let X and Y be Banach spaces and T : X → Y a bounded linear operator. Then
the image of T is closed in Y if and only if there is c > 0 such that
d(x, ker T ) ≤ ckT xk
for all x ∈ X.
Proof. Let Z be the image of T . Then the canonical map Te : X/kerT → Z induced by T is a
bounded linear isomorphism. Notice that Te(x̄) = T x for all x ∈ X, where x̄ := x+ker T ∈ X/ ker T .
Now 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. So, there is c > 0 such that d(x, ker T ) = kx̄kX/ ker T ≤
ckTe(x̄)k = ckT (x)k for all x ∈ X. So, the necessary condition follows.
For the converse, 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). So, y ∈ Z. The proof is finished.
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.
Example 9.3. Let X := {f ∈ C b (0, 1)
∩ C ∞ (0, 1)
: ∈ f0 C b (0, 1)}.
Define T : f ∈ X 7→ ∈ f0
b b
C (0, 1). Suppose that X and C (0, 1) both are equipped with the sup-norm. Then T is a closed
unbounded operator.
Proof. Note that if a sequence fn → f in X and fn0 → g in C b (0, 1). Then f 0 = g. Hence T is
closed. In fact, if we fix some 0 < c < 1, then by the Fundamental Theorem of Calculus, we have
Z x Z x
0 = lim(fn (x) − f (x)) = lim( (fn0 (t) − f 0 (t))dt) = (g(t) − f 0 (t))dt
n n c c
Rx Rx 0
for all x ∈ (0, 1). This implies that we have c g(t)dt = c f (t)dt. So g = f 0 on (0, 1).
On the other hand, since kT xn k∞ = n for all n ∈ N. Thus T is unbounded as desired.
Corollary 10.4. 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 ∗∗ . Notice that (xn ) is weakly
convergent if and only if (Qxn ) is w∗ -convergent. So, (Qxn (x∗ )) is bounded for all x∗ ∈ X ∗ . Notice
that the dual space X ∗ must be complete. So, we can apply the Uniform Boundedness Theorem
to see that (Qxn ) is bounded and so is (xn ).
19
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 11.1. Let V be an inner product space. For all x, y ∈ V , we always have:
(i): (Cauchy-Schwarz inequality): |(x, y)| ≤ kxkkyk Consequently, the inner product on
V × V is jointly continuous.
(ii): (Parallelogram law): kx + yk2 + kx − yk2 = 2kxk2 + 2kyk2
Furthermore, a norm k · k 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) = (kx + yk2 − kx − yk2 ) and Im(x, y) = (kx + iyk2 − kx − iyk2 )
4 4
for all x, y ∈ X.
Example 11.2. It follows from Proposition 11.1 immediately that `2 is a Hilbert space and `p is
not for all p ∈ [1, ∞] \ {2}.
From now on, all vector spaces are assumed to be a complex inner product spaces. Recall that
two vectors x and y in an inner product space V are said to be orthogonal if (x, y) = 0.
Proposition 11.3. (Bessel0 s inequality) : Let {e1 , ..., eN } be an orthonormal set in an inner
product space V , that is (ei , ej ) = 1 if i = j, otherwise is equal to 0. Then for any x ∈ V , we have
N
X
|(x, ei )|2 ≤ kxk2 .
i=1
Corollary 11.4. Let (ei )i∈I be an orthonormal set in an inner product space V . Then for any
element x ∈ V , the set
{i ∈ I : (ei , x) 6= 0}
is countable.
20 CHI-WAI LEUNG
Definition 11.5. A Hilbert space is a Banach space whose norm is given by an inner product.
In the rest of this section, X always denotes a complex Hilbert space with an inner product (·, ·).
Proposition 11.6. 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 ), that is, 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
Proposition 11.7. 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 ;
21
Note : there are only countable many (x, ei ) 6= 0 by Corollary 11.4, so the sums in (iii) and (iv)
are convergent by Proposition 11.6.
Remark 11.9. Recall that a vector space dimension of X is defined by the cardinality of a maximal
linearly independent set in X.
Notice that if X is finite dimensional, then the orthonormal dimension is the same as the vector
space dimension.
Also, the vector space dimension is larger than the orthornormal dimension in general since every
orthogonal set must be linearly independent.
We say that 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 x0 ) = (x, x0 ) for all x, x0 ∈ X. In this case U is called a unitary
operator.
Theorem 11.10. 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 base of X, then {U ei }i∈I is also an orthonormal base of Y .
Thus the necessary part follows from Proposition 11.8 at once.
Corollary 11.11. Every separable Hilbert space is isomorphic to `2 or Cn for some n.
Proof. Let X be a separable Hilbert space.
If dim X < ∞, then it is clear that X is isomorphic to Cn for n = dim X.
Now suppose that dim X = ∞ and its orthonormal dimension is larger √ than |N|, that is X has an
orthonormal base {fi }i∈I with |I| > |N|. Note that since kfi − fj k = 2 for all i, j ∈ I with i 6= j.
This implies that B(ei , 1/4) ∩ B(ej , 1/4) = ∅ for i 6= j.
22 CHI-WAI LEUNG
On the other hand, if we let D be a countable dense subset of X, then B(fi , 1/4) ∩ D 6= ∅ for all
i ∈ I. So for each i ∈ I, we can pick up an element xi ∈ D ∩ B(fi , 1/4). Therefore, one can define
an injection from I into D. It is absurd to the countability of D.
Proposition 12.2. Suppose that M is a closed subspace. Let u ∈ X and w ∈ M . Then the
followings are equivalent:
(i): ku − wk = d(u, M );
(ii): u − w ⊥ M , that is (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 ≤ ku − w − txk2 = ku − wk2 + ktxk2 − 2Re(u − w, tx) = d2 + ktxk2 − 2Re(u − w, tx).
This implies that
(12.1) 2Re(u − w, x) ≤ tkxk2
for all t > 0 and for all x ∈ M . So by considering −x in Eq.12.1, we obtain
2|Re(u − w, x)| ≤ tkxk2 .
for all t > 0. This implies that Re(u − w, x) = 0 for all x ∈ M . Similarly, putting ±ix into Eq.12.1,
we have Im(u − w, x) = 0. So (ii) follows.
For (ii) ⇒ (i), we need to show that ku − wk2 ≤ ku − xk2 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
ku − xk2 = k(u − w) + (w − x)k2 = ku − wk2 + kw − xk2 ≥ ku − wk2 .
23
Remark 12.6. It is worthwhile pointing out that for a general Banach space X and a closed
subspace M of X, it May Not have a complementary Closed subspace N of M , that is X = M ⊕N .
If M has a complementary closed subspace X, we say that M is complemented in X.
Example 12.7. If M is a finite dimensional subspace of a normed space X, then M is comple-
mented in X.
In fact, if M is spanned by {ei : i = 1, 2.., m}, then M is closed and by the Hahn-Banach Theorem,
for each
Tmi = 1, ..., m, there is e∗i ∈ X ∗ such that e∗i (ej ) = 1 if i = j, otherwise, it is equal to 0. Put
∗
N := i=1 ker ei . Then X = M ⊕ N .
Theorem 12.9. 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 kf k = kvf k.
P
Furthermore, if (ei )i∈I is an orthonormal base 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. So z − vf = 0.
Now for proving the existence of vf , it suffices to show the case kf k = 1. Then ker f is a closed
proper subspace. Then by the orthogonal decomposition again, we have
X = ker f ⊕ (ker f )⊥ .
Since f 6= 0, we have (ker f )⊥ is linear isomorphic to C. Also note that the restriction of f
on (ker f )⊥ is of norm one. Hence there is an element vf ∈ (ker f )⊥ with kvf k = 1 such that
f (vf ) = kf |(ker f )⊥ k = 1 and (ker f )⊥ = Cvf . So for each element x ∈ X, we have x = z + αvf for
24 CHI-WAI LEUNG
Corollary 12.10. With the notation as in Theorem 12.9, Define the map
(12.2) Φ : f ∈ X ∗ 7→ vf ∈ X, i.e., f (y) = (x, Φ(f ))
for all y ∈ X and f ∈ X ∗ .
And if we define (f, g)X ∗ := (vg , vf )X for f, g ∈ X ∗ . Then (X ∗ , (·, ·)X ∗ ) becomes a Hilbert space.
Moreover, Φ is an anti-unitary operator from X ∗ onto X, that is 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.12.2 above.
Theorem 12.12. 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 7.9, (jM (xn )) has a w∗ -convergent subsequence
w∗
(jM (xnk )). Put jM (xnk ) −−→ f ∈ M ∗ , that is jM (xnk )(z) → f (z) for all z ∈ M . The Riesz
Representation will assure that there is a unique element m ∈ M such that jM (m) = f . So we
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 . So the dual space M ∗ is also separable (see Proposition5.7). Let
Q : M −→ M ∗∗ be the natural canonical mapping. To apply the Banach’s result Proposition 7.9
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. So 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 as desired.
25
Remark 12.13. It is well known that we have the following Theorem due to R. C. James (the
proof is highly non-trivial):
A normed space X is reflexive if and only if every bounded sequence in X has a weakly convergent
subsequence.
Theorem 12.12 can be obtained by the James’s Theorem directly. However, Theorem 12.12 gives a
simple proof in the Hilbert spaces case.
Remark 13.1. 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
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 . So, 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 13.3. Let T ∈ B(X). Then there is a unique element T ∗ in B(X) such that
(13.2) (T x, y) = (x, T ∗ y)
In this case, T ∗ is called the adjoint operator of T .
26 CHI-WAI LEUNG
Proof. We first show the uniqueness. Suppose that there are S1 , S2 in B(X) which satisfy the
Eq.13.2. Then (x, S1 y) = (x, S2 y) for all x, y ∈ X. Eq.13.1 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 ∗ . The Riesz Representation Theorem implies that there is a
unique element y ∗ ∈ X such that (T x, y) = (x, y ∗ ) for all x ∈ X and kfy k = ky ∗ k. On the other
hand, we have
|fy (x)| = |(T x, y)| ≤ kT kkxkkyk
for all x, y ∈ X and thus kfy k ≤ kT kkyk. If we put T ∗ (y) := y ∗ , then T ∗ satisfies the Eq.13.2.
Also, we have kT ∗ yk = ky ∗ k = kfy k ≤ kT kkyk for all y ∈ X. So T ∗ ∈ B(X) with kT ∗ k ≤ kT k
indeed. Hence T ∗ is as desired.
Example 13.7. If X = Cn and D = (aij )n×n an n × n matrix, then D∗ = (aji )n×n . In fact, notice
that
aji = (Dei , ej ) = (ei , D∗ ej ) = (D∗ ej , ei ).
So if we put D∗ = (dij )n×n , then dij = (D∗ ej , ei ) = aji .
∞
P∞ X
Example 13.8. Let `2 (N) := {x : N → C : 2
i=0 |x(i)| < ∞}. And put (x, y) := x(i)y(i).
i=0
Define the operator D ∈ B(`2 (N)) (called the unilateral shift) by
Dx(i) = x(i − 1)
for i ∈ N and where we set x(−1) := 0, that is 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, .., that is D∗ (x(0), x(1), ...) = (x(1), x(2), ....).
Indeed one 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 13.9. Let `∞ (N) = {x : N → C : supi≥0 |x(i)| < ∞} and kxk∞ := supi≥0 |x(i)|. For
each x ∈ `∞ , define Mx ∈ B(`2 (N)) by
Mx (ξ) := x · ξ
for ξ ∈ `2 (N),
where (x · ξ)(i) := x(i)ξ(i); i ∈ N.
Then kMx k = kxk∞ and Mx∗ = Mx , where x(i) := x(i).
Definition 13.10. 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.
Proposition 13.14. Let (E, k · k) be a Banach space. Let M and N be the closed subspaces of E
such that
E =M ⊕N . . . . . . . . . . . . (∗)
Define an operator Q : E −→ E by Q(y + z) = y for y ∈ M and z ∈ N . Then Q is bounded. In
this case, Q is called the projection with respect to the decomposition (∗).
Furthermore, if E is a Hilbert space, then N = M ⊥ (and hence (∗) is the orthogonal decomposition
of E with respect to M ) if and only if Q satisfies the conditions: Q2 = Q and Q∗ = Q. And Q is
called the orthogonal projection (or projection for simply) with respect to M .
Proof. For showing the boundedness of Q, by using the Closed Graph Theorem, we need to show
that if (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 and u = v ⊕ w, where yn , v ∈ M and zn , w ∈ N , then Qxn = yn .
Notice that (zn ) is a convergent sequence in E 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 M and N are closed, we have u ∈ M and z ∈ N . Therefore, we have Qx = u as desired.
So Q∗ = Q.
The converse of the last statement follows from Proposition 13.13 at once because ker Q = N and
imQ = M .
The proof is complete.
Proposition 13.15. 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. It first follows from Proposition 13.14 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 because M = ker(I − Q) and I − Q is bounded. Also it is clear that
X = Q(X) ⊕ (I − Q)X with ker Q = M ⊥ . Hence M is the corresponding closed subspace of X,
that is M ∈ M and PM = Q as desired.
For the final assertion, Part (i) and (ii) follow immediately from the orthogonal decompositions
X = M ⊕M ⊥ = N ⊕N ⊥ and together with the clear facts that M ⊆ N if and only if N ⊥ ⊆ M ⊥ .
Example 14.3. 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 14.4. Let E = (c00 (N), k · k∞ ) (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, kT xk∞ ≤ kxk∞ 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
30 CHI-WAI LEUNG
From this we see that σp (T ) = {1, 21 , 13 , ...}. And if λ ∈ / {1, 21 , 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
So, (T − λ)−1 is unbounded if λ = 0 and thus 0 ∈ σ(T ).
On the other hand, if λ 6= 0, then (T − λ)−1 is bounded. In fact, if λ = a + ib 6= 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
k(T − λ)−1 k = sup |( − λ)−1 | < η −1 < ∞.
k∈N k + 1
It can now be concluded that σ(T ) = {1, 21 , 13 , ...} ∪ {0}.
Example 14.7. Let E = `2 (N) and D ∈ B(E) be the right unilateral shift operator as in Example13.8.
Recall that Dx(k) := x(k − 1) for i ∈ 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 ∈ `2 (N) satisfy the equation Dx = λx. Then by the definition of D, we
have
x(k − 1) = λx(k) · · · · · · · · · (∗)
31
for all k ∈ N.
If λ 6= 0, then we have x(k) = λ−1 xk−1 for all i ∈ N. Since x(−1) = 0, this forces x(k) = 0 for all
i, that is x = 0 in `2 (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, kDk = 1. Proposition 14.5 tells us that
σ(D) ⊆ {λ ∈ C : |λ| ≤ 1}.
Notice 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 when |λ| = 1 first.
Let e1 = (1, 0, 0, ...) ∈ `2 (N). Then by the assumption, there is x ∈ `2 (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 since
e1 (0) = 1 and e1 (k) = 0 for all k ≥ 1. Also since |λ| = 1, it turns out that |x(0)| = |x(k)| for all
k ≥ 1. As x ∈ `2 (N), this forces x = 0. However, it is absurd because Dx = λx + e1 .
Now we consider the case when |λ| < 1.
Notice that by Proposition 13.13, we have
⊥
im(D − λ) = ker(D − λ)∗ = ker(D∗ − λ).
Thus if D − λ is surjective, we have ker(D∗ − λ) = (0) and hence λ ∈ / σp (D∗ ).
Notice that the adjoint D∗ of D is given by the left shift operator, that is,
D∗ x(k) = x(k + 1) · · · · · · · · · (∗∗)
for all k ∈ N.
Now when D∗ x = µx for some µ ∈ C and x ∈ `2 (N), by using Eq.(∗∗), which is equivalent to saying
that
x(k + 1) = µx(k)
k
for all k ∈ N. So as |λ| = |λ| < 1, if we set x(0) = 1 and x(k + 1) = λ x(0) for all k ≥ 1, then
x ∈ `2 (N) and D∗ x = λx. Hence λ ∈ σp (D∗ ) which leads to a contradiction.
The proof is finished.
Lemma 15.1. 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 kT xk ≥ ckxk for all x ∈ H.
Proof. The necessary part is clear.
Now we are going to show the converse. We first to show the case when T is selfadjoint. It is clear
that T is injective from the assumption. So by the Open Mapping Theorem, it remains 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. So (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.
32 CHI-WAI LEUNG
Definition 15.2. Let T ∈ B(X). We say that T is positive, write T ≥ 0, if (T x, x) ≥ 0 for all
x ∈ H.
Remark 15.3. It is clear that a positive operator is selfadjoint by Proposition 13.12 at once.
In particular, all projections are positive.
Remark 15.5. In Proposition 15.4, 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
So by Lemma 15.1 we have shown that T − M is not invertible and hence M ∈ σ(T ) if T ≥ 0.
Now for any selfadjoint operator T if we consider T − m, then T − m ≥ 0. Thus we have M − m =
M (T − m) ∈ σ(T − m) by the previous case. It is clear that σ(T − c) = σ(T ) − c for all c ∈ C.
Therefore we have M ∈ σ(T ) for any self-adjoint operator.
We are now claiming that m(T ) ∈ σ(T ). Notice that M (−T ) = −m(T ). So we have −m(T ) ∈
σ(−T ). It is clear that σ(−T ) = −σ(T ). Then m(T ) ∈ σ(T ).
Finally, we are going to show σ(T ) ⊆ [m, M ].
Indeed, since T − m ≥ 0, then by Proposition 15.4, we have σ(T ) − m = σ(T − m) ⊆ [0, ∞). This
gives σ(T ) ⊆ [m, ∞).
On the other hand, similarly, we consider M − T ≥ 0. Then we get M − σ(T ) = σ(M − T ) ⊆ [0, ∞).
This implies that σ(T ) ⊆ (−∞, M ]. The proof is finished.
Definition 16.1. 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 16.2. Let U be the closed unit ball of H. It is clear that 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.
Also it is clear that 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 equivalent to it is compact.
Example 16.3. The identity operator I : H → H is compact if and only if dim H < ∞.
Example 16.4. Let H = `2 ({1, 2...}). Define T x(k) := x(k)k for k = 1, 2.... Then T is compact.
In fact, if we let (xn ) be a bounded sequence in `2 , then by the diagonal argument, we can find
a subsequence ym := T xm of T xn such that lim ym (k) = y(k) exists for all k = 1, 2... Let
m→∞
L := supn kxn k22 . Since |ym (k)|2 ≤ kL2 for all m, k, we have y ∈ `2 . Now let ε > 0. Then one can
find a positive integer N such that k≥N 4L/k 2 < ε. So we have
P
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
Theorem 16.5. 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.
34 CHI-WAI LEUNG
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. So 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 not. Then by the compactness of T again, we can find a subsequence (xi ) of (xn ) such
that T xi converges to y 0 with y 6= y 0 . Thus there is z ∈ H such that (y, z) 6= (y 0 , z). On the other
hand, if we let x be the weakly limit of (xn ), then (xn , w) → (x, w) for all w ∈ H. So we have
(y, z) = lim(T xj , z) = lim(xj , T ∗ (z)) = (x, T ∗ z) = (T x, z).
j j
Proposition 16.8. Let T ∈ K(H) and let c ∈ C with c 6= 0. Then T − c has a closed range.
1 1
Proof. Notice that since T ∈ K(H), so if we consider T − I, we may assume that c = 1.
c c
Let S = T − I. 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 not. By considering a subsequence
zn
of (zn ), we may assume that we may assume that kzn k → ∞. Put vn := ∈ (ker S)⊥ .
kzn k
35
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)⊥ . Also from this we have Svn → Sw. On the other hand,
we have Sw = limn Svn = limn T vn − limn vn = w − w = 0. So w ∈ ker S ∩ (ker S)⊥ . It follows
that w = 0. However, since vn → w and kvn k = 1 for all n. It leads to a contradiction. So (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 finished.
Theorem 16.9. Fredholm Alternative Theorem : Let T ∈ K(H)sa and let 0 6= λ ∈ C. Then
T − λ is injective if and only if T − λ is surjective.
Proof. Since T is selfadjoint, σ(T ) ⊆ R. So if λ ∈ C \ R, then T − λ is invertible. So the result
holds automatically.
Now consider the case λ ∈ R \ {0}.
Then T − λ is also selfadjoint. From this and Proposition 13.13, we have ker(T − λ) = (im(T − λ))⊥
and (ker(T − λ))⊥ = im(T − λ).
So the proof is finished by using Proposition 16.8 immediately.
Corollary 16.10. 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 15.6 are non-zero, then both are the
eigenvalues of T and kT k = max |λ|.
λ∈σp (T )
Proof. It follows from the Fredholm Alternative Theorem at once. This together with Theorem
15.6 imply the last assertion.
Example 16.11. Let T ∈ B(`2 ) be defined as in Example 16.4. We have shown that T ∈ K(`2 )
and it is clear that T is selfadjoint. Then by Corollary 16.10 and Corollary 16.7, we see that
σ(T ) = {0, 1, 12 , 13 , .....}.
Lemma 16.12. 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. If we fix µ ∈ σ(T ) \ {0} and put Iµ := {λ ∈ σ(T ) : |λ| = |µ|},
then we have M
dim Eλ < ∞.
λ∈Iµ
Proof. We first notice that dim Eλ < ∞ for all λ ∈ σp (T ) \ {0} because the restriction T |Eλ is also
a compact operator on Eλ .
0 0
On theLother hand, since T is selfajoint, we also have Eλ ⊥Eλ0 for λ, λ ∈ σp (T ) with λ 6= λ . Let
V := λ∈Iµ Eλ . Suppose that dim V = ∞. Then |Iµ | = ∞. So, we can find an infinite sequence in
Iµ such that λm 6= λn for m 6= n. Now choose vn ∈ Eλn with kvn k = 1 for each λn . Then vn ⊥vm
for n 6= m. This implies that kT vn − T vm k2 = |λn |2 + |λm |2 = 2|µ|2 > 0 for m 6= n. So (T vn ) has
no convergent subsequences which contradicts to T being compact.
Theorem 16.13. Let T ∈ K(H)sa . And suppose that dim H = ∞. Then σ(T ) = {λ1 , λ2 , ....}∪{0},
where (λn ) is a sequence of real numbers with λn 6= λm for m 6= n and |λn | ↓ 0.
36 CHI-WAI LEUNG
Proof. Note that since kT k = max(|M (T )|, |m(T )|) and σ(T ) \ {0} = σp (T ) \ {0}. So by Corollary
16.10, there is |λ1 | = max |λ| = kT k. Since dim Eλ1 < ∞, then Eλ⊥1 6= 0. Then by considering
λ∈σp (T )
the restriction of T2 := T |Eλ⊥1 6= 0, there is |λ2 | = maxλ∈σp (T2 ) |λ| = kT2 k. Notice that λ2 ∈ σp (T )
and |λ2 | ≤ |λ1 | because kT2 k ≤ kT k. To repeat the same step, we can get a sequence (λn ) such
that (|λn |) is decreasing.
Now we claim that limn |λn | = 0.
Otherwise, there is η > 0 such that |λn | ≥ η for all n. If we let vn ∈ Eλn with kvn k = 1 for all n.
Notice 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 kT wn − T wn k2 = kvn − vm k2 = 2 for m 6= n.
This is a contradiction since T is compact. So limn |λn | = 0.
Finally we need to check σ(T ) = {λ1 , λ2 , ...} ∪ {0}.
In fact, let µ ∈ σp (T ). Since |λn | ↓ 0, we can find a subsequence n1 < n2 < .... of positive integers
such that
|λ1 | = ... = |λn1 | > |λn1 +1 | = ... = |λn2 | > |λn2 +1 | = .... = |λn3 | > |λn3 +1 | = ....
Then we can choose N such that |λnN +1 | < |µ| ≤ |λnN |. Notice that by the construction of λn ’s
implies µ = λj for some nN −1 + 1 ≤ j ≤ nN .
The proof is finished.
Theorem 16.14. Let T ∈ K(H)sa and let (λn ) be given as in Theorem 16.13. For each λ ∈
σp (T ) \ {0}, put d(λ) := dim Eλ < ∞. Let {eλ,i : i = 1, ..., d(λ)} be an orthonormal base for Eλ .
Then we have the following orthogonal decomposition:
∞
M
(16.1) H = ker T ⊕ Eλn .
n=1
It is absurd because µ 6= λ1i for all i. So T |E ⊥ = 0 and hence E ⊥ ⊆ ker T . So we have the
decomposition (16.1). And from this we see that the family B forms an orthonormal base of
(ker T )⊥ . On the other, we have (ker T )⊥ = imT ∗ = imT . Therefore, B is an orthonormal base for
T (H) as desired.
For the last assertion,Pit needs to show that the series ∞
P
n=1 λn Pn converges to T in norm. Notice
m
that if we put Sm := n=1 λn Pn , then by the decomposition (16.1), lim Sm x = T x for all x ∈ H.
m→∞
So it suffices to show that (Sm )∞
m=1 is a Cauchy sequence in B(H). In fact we have
References
[1] J.B. Conway. A course in functional analysis, second edition, Springer-Verlag, (1990).
[2] J. Diestel, Sequences and series in Banach spaces, Springer-Verlag, (1984).
[3] H. Royden and P. Fitzpatrick, Real analysis, fourth edition, Pearson, (2010).
[4] W. Rudin. Functional analysis, second edition, McGraw-Hill Inc, (1991).
(Chi-Wai Leung) Department of Mathematics, The Chinese University of Hong Kong, Shatin, Hong
Kong
E-mail address: [email protected]