Functional Analysis2
Functional Analysis2
PIOTR HAJLASZ
(1) hx, xi ≥ 0;
(2) hx, xi = 0 if and only if x = 0;
(3) hαx, yi = αhx, yi;
(4) hx1 + x2 , yi = hx1 , yi + hx2 , yi;
(5) hx, yi = hy, xi,
Proof. The properties kαxk = |α|kxk and kxk = 0 if and only if x = 0 are
obvious. To prove the last property we need to apply the Schwarz inequality.
kx + yk2 = hx + y, x + yi = hx, xi + 2 re hx, yi + hy, yi
≤ kxk2 + 2kxk kyk + kyk2 = (kxk + kyk)2 .
(1) If K = R, then
1
hx, yi = (kx + yk2 − kx − yk2 ) ,
4
for all x, y ∈ X.
(2) If K = C, then
1 1
kx + yk2 − kx − yk2 − i kix + yk2 − kix − yk2
hx, yi =
4 4
for all x, y ∈ X.
Now the ellipsoid E = L(B n (0, 1)) is the unit ball for the inner product
n n n
X X X ai bi
ai vi , bi vi = .
i=1 i=1 i=1
λ2i
We will see later (Corollary 5.12) any real inner product space space H
of dimension n is isometrically isomorphic to `2n , i.e. Rn with the standard
inner product, so if L : `2n → H is this isometric isomorphism, the unit ball
in H is L(B n (0, 1)), so it is an ellipsoid. Thus we proved.
Theorem 1.5. A convex set in Rn is a unit ball for a norm associated with
an inner product if and only if it is an ellipsoid.
7. c1 , the space of all (complex, real) convergent sequences with the norm
k · k∞ is a Banach space.
8. c0 , the space of all (complex, real) sequences that converge to zero with
the norm k · k∞ is a Banach space.
9. Note that c0 ⊂ c ⊂ `∞ and both c0 and c are closed linear subspaces of
`∞ with respect to the metric generated by the norm.
Exercise. Prove that `∞ , c and c0 are Banach spaces.
Exercise. Prove that the spaces c and c0 are separable, while `∞ is not.
10. `p , 1 ≤ p < ∞ is the space of all (complex, real) sequences x = (xn )∞
n=1
such that
∞
!1/p
X
p
kxkp = |xn | .
n=1
It follows from the Minkowski inequality for sequences that k · kp is a norm
and that `p is a linear space. We will prove now that `p is a Banach space,
i.e. that it is complete. Let xn = (ani )∞ p
i=1 be a Cauchy sequence in ` , i.e. for
every ε > 0 there is N such that for all n, m > N
∞
!1/p
X
n m p
kxn − xm kp = |ai − ai | < ε.
i=1
Hence for each i the sequence (ani )∞
n=1 is a Cauchy sequence in K (= C or
n
R). Let ai = limn→∞ ai . Fix an integer k. Then for n, m > N we have
k
X
|ani − am p
i | <ε
p
i=1
and passing to the limit as m → ∞ yields
k
X
|ani − ai |p ≤ εp .
i=1
Now taking the limit as k → ∞ we obtain
∞
X
|ani − ai |p ≤ εp ,
i=1
i.e.
kxn − xm kp ≤ ε where x = (ai )∞
i=1 .
p p
This proves that x ∈ ` and xn → x in ` . The proof is complete.
In particular the space `2 is a Hilbert space because its norm is associated
with the inner product
∞
X
hx, yi = xi yi .
i=1
FUNCTIONAL ANALYSIS 7
11. We will prove that `p for p 6= 2 is not an inner product space. Let
x = (1, 0, 0, . . .), y = (0, 1, 0, 0, . . .). If 1 ≤ p < ∞, then
and thence the Parallelogram Law is violated. The same example can also
be used in the case p = ∞. In the real case this result can also be seen as a
consequence of the fact that the two dimensional section of the unit ball in
`p , p 6= 2, along the space generated by the first two coordinates is not an
ellipse.
Also L∞ (µ) is a Banach space with the norm being the essential supremum
of |f |. For the proofs see the notes from Analysis I.
For p = 2 the space L2 (µ) is a Hilbert space with respect to the inner
product
Z
hf, gi = f g dµ .
X
kf k = sup |f (x)|
x∈X
15. Let H 2 be the class of all holomorphic functions on the unit disc D =
{z ∈ C : |z| < 1} such that
Z 2π
1 1/2
kf kH 2 = sup |f (reiθ )|2 dθ < ∞.
0<r<1 2π 0
This space is called Hardy space H 2 . We will prove now that it is a Hilbert
space and we will find an explicit formula for the inner product.
8 PIOTR HAJLASZ
2. Linear operators
(1) I is continuous;
(2) L is continuous at 0;
(3) L is bounded.
Proof. The implication (1) ⇒ (2) is obvious. (2) ⇒ (3) For if not
there would exist a sequence xn ∈ X such that kLxn k > nkxn k. Then
kL(xn /(nkxn k))k > 1, but on the other hand kL(xn /(nkxn k))k → 0, be-
cause xn /(nkxn k) → 0 which is an obvious contradiction. (3) ⇒ (1) Let
xn → x. Then Lxn → Lx. Indeed,
kLx − Lxn k = kL(x − xn )k ≤ Ckx − xn k → 0 .
FUNCTIONAL ANALYSIS 9
Then
kLxk ≤ kLk kxk for all x ∈ X.
Indeed,
x
x
kLxk =
L kxk
= kxk
L
≤ kxk kLk .
kxk kxk
Thus L is bounded if and only if kLk < ∞. Moreover kLk is the smallest
constant C for which the inequality
kLxk ≤ Ckxk for all x ∈ X
is satisfied. kLk is called the operator norm.
The class of bounded operators L : X → Y is denoted by B(X, Y ). We
also write B(X) = B(X, X). Clearly B(X, Y ) has a structure of a linear
space.
Lemma 2.2. B(X, Y ) equipped with the operator norm is a normed space.
Proof. Indeed, K is a Banach space and the result follows from Theo-
rem 2.3 2
Definition. We say that the two normed spaces X and Y are isomorphic
if there is an algebraic isomorphism of linear spaces L ∈ B(X, Y ) (i.e. it
is one-to-one and surjection) such that L−1 ∈ B(Y, X). The mapping L is
called an isomorphism of normed spaces X and Y . If X and Y are Banach
spaces we call it isomorphism of Banach spaces X and Y .
Proof. It easily follows from the fact that continuity of a linear mapping
is equivalent with its boundedness. 2
defines a bounded functional x∗ ∈ c∗0 with kx∗ k = ksk1 . On the other hand
if x∗ ∈ c∗0 , then there is unique x ∈ `1 such that x∗ can be represented by
(2.4). This proves that the space c∗0 is isometrically isomorphic to `1 .
Hence
k
X k
X
kx∗ k ≥ hx∗ , z k i = zi hx∗ , ei i = |si | .
| {z }
i=1 si i=1
Letting k → ∞ we have
∞
X
ksk1 = |si | ≤ kx∗ k
i=1
`1
which proves that s ∈ along with the estimate (2.6). Now it easily follows
that x∗ satisfies (2.4). Indeed, if x = (xi ) ∈ c0 and
k
X
xk = (x1 , . . . , xk , 0, 0, . . .) = xi ei ,
i=1
Pk P∞
then xk
→ x in c0 and i=1 xi si → i=1 xi si , because s ∈ `1 . Hence passing
to the limit in the equality
k
X
hx∗ , xk i = xi si
i=1
yields (2.4). The proof is complete. 2
Exercise. Prove that the dual space c∗ is isometrically isomorphic to `1 .
Theorem 2.11. If s = (si ) ∈ `∞ , then
∞
X
(2.7) hx∗ , xi = si xi for x = (xi ) ∈ `1
i=1
defines a bounded functional x∗ ∈ (`1 )∗ with kx∗ k = ksk∞ . On the other hand
if x∗ ∈ (`1 )∗ , then there is unique s ∈ `∞ such that x∗ can be represented by
(2.7). This proves that the space (`1 )∗ is isometrically isomorphic to `∞ .
Proof. The proof is pretty similar to the previous one, so we will be short.
If s ∈ `∞ , then it is easily seen that x∗ given by (2.4) defines a functional
on `1 with kx∗ k ≤ ksk∞ . Now let x∗ ∈ (`1 )∗ . It remains to prove that
there is s ∈ `∞ such that x∗ satisfied (2.7) and ksk∞ ≤ kx∗ k (uniqueness is
obvious). Let ei = (0, . . . , 0, 1, 0, . . .) ∈ `1 and si = hx∗ , ei i. Let zi = si /|si |
if si 6= 0 and zi = 0 if si = 0. Put z i = zi ei ∈ `1 , so kz i ki ≤ 1. Then
kx∗ k ≥ hx∗ , z i i = |si |. Now taking supremum over all i yields ksk∞ ≤ kx∗ k
and the result easily follows. 2
Exercise. Prove that for every s = (si ) ∈ `1 , hx∗ , xi = ∞
P
i=1 si xi defines a
bounded functional on `∞ with kx∗ k = ksk1 .
Later we will see that not every functional in (`∞ )∗ can be represented by
an element of `1 , and the above exercise proves only that `1 is isometrically
isomorphic to a closed subspace of (`∞ )∗ .
14 PIOTR HAJLASZ
Theorem 2.12. Let 1 < p < ∞. If s = (si ) ∈ `q , where q = p/(p − 1), then
∞
X
(2.8) hx∗ , xi = si xi for x = (xi ) ∈ `p
i=1
defines a bounded functional x∗ ∈ (`p )∗ with kx∗ k = kskq . On the other hand
if x∗ = (`p )∗ , then there is unique s ∈ `q such that x∗ can be represented by
(2.8). This proves that the space (`p )∗ is isometrically isomorphic to `q .
Proof. Again, the proof is very similar to those presented above, so we will
explain only the step where a tiny difference in the argument occurs. Let
x∗ ∈ (`p )∗ . The crucial point is to show that for si = hx∗ , ei i we have kskq ≤
kx∗ kp . To prove this we take zi = si |si |q−2 and z k = (z1 , . . . , zk , 0, 0, . . .) ∈ `p .
Then
k
X
∗ k ∗ k
(2.9) kx k kz kp ≥ hx , z i = |si |q .
i=1
Since kz k kp = ( ki=1 |si |q )1/p inequality (2.9) yields kx∗ k ≥ ( ki=1 |si |q )1/q
P P
and the claim follows after passing to the limit as k → ∞. 2
The last two results are special cases of the following deep result whose
proof based on the Radon-Nikodym theorem is presented in notes from Anal-
ysis I and also will be proved in Section 5.5.
Theorem 2.13. If µ is a σ-finite measure on X and 1 ≤ p < ∞, 1 < q ≤ ∞,
p−1 + q −1 = 1, then for every function g ∈ Lq (µ),
Z
(2.10) Λf = f g dµ for f ∈ Lp (µ)
X
Theorem 3.1. In a finitely dimensional linear space any two norms are
equivalent. In particular every finitely dimensional normed space is a Banach
space.
Note that the space X with respect to the norm k · k0 is locally compact and
complete. It suffices to prove that every norm in X is equivalent with k · k0 .
Let k · k be an arbitrary norm in X. Then
X n
X n n
X
0
kxk =
xi ei
≤ |xi |kei k ≤ kxk kei k = Ckxk0 .
i=1 i=1 i=1
is true for any locally integrable f . It is also clear that for every ball there is
a constant C such that (3.1) is satisfied, however, we want to find a constant
C that will be good for all the balls at the same time.
Proof. Let k be the degree of P . The space P k of polynomials of degree
less than or equal k has finite dimension, so all norms in that space are
equivalent. In particular there is a constant C > 0 such that
Z
(3.2) sup |Q| ≤ C|B(0, 1)|−1 |Q| for all Q ∈ P k .
B(0,1) B(0,1)
Proof. The unit ball with respect to the norm k · k is a convex symmetric
body K (symmetric means that x ∈ K ⇒ −x ∈ K) and any ellipsoid E
centered at 0 is a unit balls for a Hilbert norm k · k0 (we choose the inner
product so that the semi-axes are orthonrmal, see Section 1.1, Example √ 5).
Now it suffices to prove that there is an ellipsoid E such that E ⊂ K ⊂ n E.
Indeed, E ⊂ K means that kxk0 √ ≤ 1 ⇒ kxk ≤ 1 which easily implies√
kxk ≤ kxk0 for any x ∈ X and√K ⊂ n E means that kxk ≤ 1 ⇒ kxk0 ≤ n
which easily implies kxk0 ≤ n kxk for any x ∈ X. Therefore we are left
with the proof of the following result.
Theorem 3.4 (F. John ellipsoid theorem). If K is a closed convex sym-
metric body in an n-dimensional real vector space X, then there is a closed
ellipsoid E centered at 0 such that
√
E ⊂ K ⊂ nE .
√
|x1 | ≥ (1/t + 1/ n)/2 > (1/t + 1/t)/2 = 1/t (see the picture describing W ).
Now if ε is sufficiently small the ellipsoid Eε is a tiny
√ perturbation of E and
hence also the part of Eε with |x1 | ≥ (1/t + 1/ n)/2 is contained in the
interior of W . The proof is complete. 2
As an application of Theorem 3.1 we immediately obtain
Corollary 3.5. In every finitely dimensional normed space a set is compact
if and only if it is bounded and closed.
Proof. It suffices to prove that the closed unit ball centered at 0 is not
compact. To prove the lack of compactness of this ball it suffices to prove
the existence of a sequence {xi } ⊂ X such that
1
kxi k = 1, kxi − xj k ≥ for i 6= j .
2
We construct the sequence by induction. First we choose x1 with kx1 k = 1
arbitrarily. Now suppose that the elements x1 , . . . , xn have already been
defined. Let X0 = span {x1 , . . . , xn }. Since X is infinitely dimensional, X0 6=
X and hence the Riesz lemma implies that there is xn+1 ∈ X \ X0 such that
1
kxn+1 k = 1 and kxn+1 − xk ≥
2
20 PIOTR HAJLASZ
for every x ∈ X0 . 2
Note that if in the Riesz lemma dim X0 < ∞, by local compactness of X0
we can take in the proof x0 ∈ X0 such that ky0 − x0 k = % and hence the
same proof gives
Corollary 3.8. Let X0 6= X be a closed linear subspace of a normed space
X with dim X0 < ∞. Then there is y ∈ X such that
kyk = 1 and ky − xk ≥ 1 for all x0 ∈ X0 .
In general Corollary 3.8 does not hold if dim X0 = ∞ as the next example
shows, see, however, Theorem 14.11.
Example. Consider the closed linear subspace X of C[0, 1] consisting of
functions vanishing at 0. Let
n Z 1 o
X0 = f ∈ X : f (x) dx = 0 .
0
It is easy to see that X0 is a proper closed linear subspace of X. We will
prove that there is no function f ∈ X such that
(3.4) kf k = 1 and kf − gk ≥ 1 for all g ∈ X0 .
Assume that such a function f ∈ X exists. Since f is continuous, f (0) = 0
and kf k = supx∈[0,1] |f (x)| = 1, we conclude that
Z 1
(3.5) |f (x)| dx < 1 .
0
For every h ∈ X \ X0 we set
R1
f (x) dx
g = f − ch, c = R01
0 h(x) dx
and note that the denominator in nonzero, because h 6∈ X0 . Clearly g ∈ X0
and (3.4) yields
1 ≤ kf − gk = kf − (f − ch)k = |c| khk ,
i.e. Z 1 Z 1
h(x) dx ≤ f (x) dx sup |h(x)| .
0 0 x∈[0,1]
The only difference between the seminorm and the norm is that it can vanish
on nonzero elements. A seminorm vanishes on a linear subspace if X.
If X is a normed space and Y ⊂ X is a linear subspace, then we equip
X/Y with a seminorm
k[x]k = inf kzk = inf kx − yk = dist (x, Y ) .
z∈[x] y∈Y
Theorem 4.3. The quotient space X/Y is a normed space if and only if Y
is a closed subspace of X.
Then
∞
X
(4.1) kT λk = k λn xn k ≤ kλk1 ,
n=1
so T is continuous. We will prove that T is a surjection. For every x ∈ X,
every positive integer k and every ε > 0 there is n > k such that
x − kxkxn
< ε .
Indeed, it is obvious for x = 0; if x 6= 0 the inequality is equivalent to
x
ε
− xn
<
kxk kxk
and the existence of xn follows from the density in the unit sphere. Let n1
be such that
ε
kx − λn1 xn1 k < , where λn1 = kxk .
2
Let n2 > n1 be such that
ε ε
k(x − λn1 xn1 ) − λn2 xn2 k < , where λn2 = kx − λn1 xn1 k < .
4 2
Let n3 > n2 be such that
ε ε
k(x−λn1 xn1 −λn2 xn2 )−λn3 xn3 k < , where λn3 = kx−λn1 xn1 −λn2 xn2 k <
8 4
etc. We obtain a sequence λn1 , λn2 , . . . such that
ε
(4.2) λnk+1 = kx − (λn1 xn1 + . . . + λnk xnk )k < k .
2
Let
λx,ε = (λ1 , λ2 , λ3 , . . .)
where we put λi = 0 if i 6∈ {n1 , n2 , n3 , . . .}. Clearly λx,ε ∈ `1 , because
ε ε ε
kλx,ε k1 ≤ kxk + + + + . . . = kxk + ε .
2 4 8
Now continuity of T and (4.2) implies that T (λx,ε ) = x. This proves that T
is a surjection onto X. Let
Y = ker T = {λ ∈ `1 : T (λ) = 0} .
Thus T induces an algebraic isomorphism of linear spaces `1 /Y onto X.
Continuity of T implies that Y is a closed subspace of `1 , so `1 /Y is a
Banach space. Note that if T (λ) = x, then `1 /Y 3 [λ] = {γ : T (γ) = x}.
We will prove that this algebraic isomorphism is actually an isometry. To
this end we have to prove that if T (λ) = x, then k[λ]k = kxk. Let T (λ) = x.
Since T (λx,ε ) = x we have λx,ε ∈ [λ] and hence
k[λ]k ≤ kλx,ε k ≤ kxk + ε
for every ε > 0 and thus
(4.3) k[λ]k ≤ kxk .
FUNCTIONAL ANALYSIS 25
On the other hand if T (λ) = x, then for every γ ∈ [λ] inequality (4.1) implies
that kxk = kT (γ)k ≤ kγk1 and hence
(4.4) k[λ]k = inf kγk1 ≥ kxk .
γ∈[λ]
The two inequalities (4.3) and (4.4) imply that k[λ]k = kxk. This proves
that T is an isometry of `1 /Y onto X. 2
Definition. For a closed subspace M ⊂ X of a normed space X we define
the annihilator
M ⊥ = {x∗ ∈ X ∗ : hx∗ , yi = 0 for all y ∈ M } .
Proof. Let π : X → X/M , π(x) = [x] be the quotient map. We define the
map T : (X/M )∗ → X ∗ by
hT (z ∗ ), xi = hz ∗ , [x]i for z ∗ ∈ (X/M )∗ and x ∈ X.
Since
|hT (z ∗ ), xi| ≤ kz ∗ k k[x]k ≤ kz ∗ k kxk ,
T is bounded and
(4.5) kT (z ∗ )k ≤ kz ∗ k .
We actually have equality in (4.5). To see this we need to prove opposite
inequality. Given ε > 0 let [x] ∈ X/M be such that
k[x]k = 1, |hz ∗ , [x]i| ≥ kz ∗ k − ε .
Since
1 = k[x]k = inf kyk ,
y∈[x]
there is y ∈ [x] such that kyk < 1 + ε and obviously [y] = [x]. Hence
kz ∗ k − ε ≤ |hz ∗ , [x]i| = |hz ∗ , [y]i| = |hT (z ∗ ), yi|
≤ kT (z ∗ )k(1 + ε)
and letting ε → 0 yields kz ∗ k ≤ kT (z ∗ )k which together with (4.5) proves
kT (z ∗ )k = kz ∗ k .
We proved that T is an isometric embedding of (X/M )∗ onto a closed sub-
space of X ∗ . Actually
T ((X/M )∗ ) ⊂ M ⊥ .
2Even if X is only a normed space.
26 PIOTR HAJLASZ
L̃ : X/ ker L → Y, L̃([x]) = Lx
is bounded, kL̃k ≤ kLk and it is a bijection onto L̃(X) = L(X). Now the
composition
L̃ η
X/ ker L −→ Y −→ Y /V
is a bijection, so it is an isomorphism of Banach spaces. Hence
L̃ ◦ (η ◦ L̃)−1 : Y /V → L(X)
and hence
∞
X 1/2
k(xi )k = kxi k2i .
i=1
5. Hilbert spaces
Since ϕ ∈ (`1 )∗
W = {x ∈ `1 : ϕ(x) = 1}
is convex and closed. If y = (1, 1, 1, . . .), then
∞
X
W = {x ∈ `1 : xi = 1}
n=1
so for x ∈ W , kxk = ∞
P
n=1 |xn | ≥ 1 and hence every vector x ∈ W with
xn ≥ 0 for all n has the smallest norm equal 1. Thus we have infinitely many
vectors of smallest norm.
On the other hand if y = (1/2, 2/3, 3/4, 4/5, . . .), then
∞
n X n o
W = x ∈ `1 : xn = 1 .
n+1
n=1
30 PIOTR HAJLASZ
Observe that this is a very different situation than in the case of Theo-
rem 4.6.
Proof of Theorem 5.2. (a) Uniqueness is easy. If
x = x0 + y 0 = x00 + y 00 , x0 , x00 ∈ M, y 0 , y 00 ∈ M ⊥ ,
then
M 3 x0 − x00 = y 00 − y 0 ∈ M ⊥ .
Since M ∩ M ⊥ = {0} (because hx, xi = 0 implies x = 0) we conclude that
x0 = x00 and y 0 = y 00 . Thus we are left with the proof of the existence of the
decomposition. The set
x + M = {x + y : y ∈ M }
is convex and closed. Let Qx be the element of the smallest norm in x + M
and let P x = x − Qx. Clearly x = P x + Qx. Since Qx ∈ x + M , it follows
that P x ∈ M . We still need to prove that Qx ∈ M ⊥ . To this end we have to
prove that hQx, yi = 0 for all y ∈ M . We can assume that kyk = 1. Denote
z = Qx. The minimizing property of Qx shows that
hz, zi = kzk2 ≤ kz − αyk2 = hz − αy, z − αyi
for all α ∈ K. Hence
hz, zi ≤ hz, zi + |α|2 hy, yi −αhy, zi − αhz, yi ,
| {z }
1
(b) If x ∈ H, and X
sF (x) = x̂(α)uα ,
α∈F
then
(5.3) kx − sF (x)k ≤ kx − sk
for every s ∈ MF with the equality only for s − sF (x). Moreover
X
(5.4) |x̂(α)|2 ≤ kxk2 .
α∈F
Proof. Part (a) is obvious. To prove (b) denote sF = sF (x) and observe
that sˆF (α) = x̂(α) for α ∈ F . That means (x − s + F ) ⊥ MF . In particular
hx − sF , sF − si = 0 and hence
(5.5) kx − sk2 = k(x − sF ) + (sF − s)k2 = kx − sF k2 + ksF − sk2
which implies (5.3). Now (5.5) with s = 0 gives
kxk2 = ks − sF k2 + ksF k2 ≥ ksF k2
which is (5.4). 2
Remark. The part (b) says that sF (x) is the best unique approximation
of x in MF , i.e. (see Theorem 5.2) sF (x) is the orthogonal projection of x
onto MF .
If A is any set, then we define `∞ (A) to be the Banach space of all bounded
functions on A (no measurability condition), and `p (A), 1 ≤ p < ∞ is the
Banach space of p-integrable functions with respect to the counting measure.
Thus ϕ ∈ `p (A), 1 ≤ p < ∞ if ϕ(α) 6= 0 for at most countably many α and
X X
|ϕ(α)|p = |ϕ(α)|p < ∞ .
α∈A α∈A
ϕ(α)6=0
Note that X X
|ϕ(α)|p = sup |ϕ(α)|p
α∈A F α∈F
(b) f : X → Y is continuous;
(c) X has a dense subset X0 on which f is an isometry;
(d) f (X0 ) is dense in Y .
⊥
Proof. (a) ⇒ (b). Suppose P is not dense, i.e. P 6= H. Let u ∈ P ,
kuk = 1. Then the set {u} ∪ {uα }α∈A is orthonormal which contradicts
maximality of {uα }α∈A .
(b) ⇒ (c). It follows from the previous theorem
ˆ: P = H → `2 (A)
is an isometry, so the Parseval identity follows.
(c) ⇒ (d) this implication follows from the Polarization identity, Proposi-
tion 1.3.
(d) ⇒ (a). If (a) were false, then there would be 0 6= u ∈ H such that
hu, uα i = 0 for all α ∈ A. Let x = y = u. We have
X
0 < hx, yi = x̂(α) ŷ(α) = 0
|{z} |{z}
α∈A 0 0
which is a contradiction. 2
Definition. Any maximal orthonormal set in H is called an orthonormal
basis.
Therefore the above theorem provides several equivalent conditions for an
orthonormal set to be a basis.
A direct application of the Hausdorff maximality theorem (equivalent to
the axiom of choice) gives
Theorem 5.10. Every orthonormal set is contained in an orthonormal ba-
sis.
P
Proof. As we know kx − |n|≤k x̂(n)en k equals to the distance of x to
the space span{en }|n|≤k (see Theorem 5.6(b)). Since span {en }n∈Z is a dense
subset of H this distance converges to 0 as k → ∞. 2
As a variant of Theorems 5.9 and 5.13 we obtain
Theorem 5.14. Let H1 , H2 , . . . be closed subspaces of a Hilbert space H such
that Hi ⊥ Hj for i 6= j and linear combinations of elements of subspaces
Hi are dense in H. Then H is isometrically isomorphic to the direct sum
of Hilbert spaces H1 ⊕ H2 ⊕ . . .. More precisely, for very x ∈ H there are
unique elements xi ∈ Hi such that
∞
X
x= xi
i=1
in the sense of convergence in H and
∞
X
2
kxk = kxi k2 .
i=1
Proof. It easily follows from the Fubini theorem that the functions hij
form an orthonormal family in L2 (µ × ν) and it remains to prove that
∞
X
kf kL2 (µ×ν) = |hf, hij i|2 for all f ∈ L2 (µ × ν) ,
i,j=1
X∞
= |hf, hij i|2 .
i,j=1
Lemma 5.20. Let W = span {tp1 , . . . , tpn } ⊂ L2 [0, 1], where −1/2 < p1 <
. . . < pn . Then for any q ≥ 0
r n
q 1 Y |q − pi |
dist L2 (t , W ) = .
2q + 1 q + pi + 1
i=1
Before we prove the lemma we will show how to complete the proof of the
Müntz theorem. First we will prove an L2 version of the Müntz theorem.
Theorem 5.21. Let −1/2 < p1 < p2 < p3 < . . . , limi→∞ pi = ∞. Then the
functions tp1 , tp2 , tp3 , . . . are linearly dense in L2 [0, 1], i.e.
(5.8) span {tp1 , tp2 , tp3 , . . .} = L2 [0, 1]
if and only if
∞
0 1
X
(5.9) =∞
pi
i=1
P0
where in the sum we omit the term with pi = 0.
by Lemma 5.22 and hence (5.10) follows. On the other hand if the sum at
(5.9) is finite, then both products
Y Y
(1 − q/pi ), (1 + (1 + q)/pi )
pi >q p1 >q
Since ∞
P
i=i0 1/(pi − 1) = ∞ we conclude from Theorem 5.21 that for any
ε > 0 there are numbers µi0 , . . . , µn such that the right hand side of (5.12)
is less than ε and hence for λ0 = λ1 = . . . = λi0 −1 = 0, λi = µi mp−1 i ,
i = i0 , . . . , n we have
sup tm − λ0 + λ1 tp1 + . . . + λn tpn < ε .
t∈[0,1]
Observe that computing the determinant using the algebraic definition that
involves permutations and then taking the common denominator gives
Pm (a1 , . . . , am , b1 , . . . , bm )
Dm = Qm .
i,j=1 (ai + bj )
Taking ai = pi , bj = pj + 1 we have
2
Q
p1 pn 1≤i<j≤n (pj − pi )
G(t , . . . , t ) = Qn
i,j=1 (pi + pj + 1)
and similarly
Qn 2
Q 2
q p1 pn i=1 (q − pi ) 1≤i<j≤n (pi − pj )
G(t , t , . . . , t ) = Qn .
+ 1)2 ni,j=1 (pi + pj
Q
(2q + 1) i=1 (q + pi + 1)
The two above formulas together with (5.13) give the result. 2
46 PIOTR HAJLASZ
Accordingly
Z
f 7→ f dν
X
FUNCTIONAL ANALYSIS 47
Since w(x) > 0 for all x, we conclude that µ(B) = 0 and hence νs ⊥ µ.8
Applying (5.15) to
f = (1 + g + . . . + g n ) χE
we obtain Z Z
(1 − g n+1 ) dν = g(1 + g + . . . + g n )w dµ .
E E
If x ∈ B, then g(x) = 1 and if x ∈ A, then g n+1 (x) decreases monotonically
to zero. Hence letting n → ∞ the left hand side converges to ν(A ∩ E) =
νa (E). On the right hand side the function that we integrate increases to a
measurable function h and the monotone convergence theorem gives
Z
νa (E) = h dµ .
E
6. Fourier Series
The C(S 1 )
and Lp (S 1 )
norms will be denoted by k·k∞ and k·kp respectively,
2 1
but the L (S ) norm will be simply denoted by k · k. As a direct application
of Theorem 5.13 we have.
Theorem 6.1. The functions en (x) = e2πinx , n ∈ Z form an orthonormal
basis in L2 (S 1 ). Hence any f ∈ L2 (S 1 ) can be represented as a series
∞
X
f= fˆ(n) en
n=−∞
where Z 1
fˆ(n) = hf, en i = f (x)e−2πinx dx ,
0
in the sense that the series converges to f in the L2 norm, i.e.
X
f − fˆ(n) en
→ 0 as k → ∞.
|n|≤k
Proof. We only need to prove that the family {en }n∈Z is an orthonormal
basis in L2 (S 1 ). It is easy to see that the functions en are orthonormal, so
we are left with the proof that linear combinations of the en ’s are dense in
L2 (S 1 ) (see Theorem 5.9(b)). Since linear combinations of the functions en
are exactly trigonometric polynomials of period 1, they are dense in C(S 1 )
by the Weierstrass theorem, and the claim follows from the density of C(S 1 )
in L2 (S 1 ). 2
FUNCTIONAL ANALYSIS 49
√ Another orthonormal
√ basis in L2 (S 1 ) is given by the functions 1,
2 cos(2πnx), 2 sin(2πnx), n = 1, 2, 3 . . . and Theorem 6.1 has an ob-
vious counterpart in this case as well. However, it is more convenient to
work with the basis ek and we will always use it in what follows.
As a first step we will derive a useful integral representation for the partial
sum of the Fourier series. Note that the partial sums of the Fourier series
are well defined for f ∈ L1 (S 1 ). We have
X X Z 1
sn (f, x) = fˆ(k)e2πikx = 2πikx
e f (y)e−2πiky dy
|k|≤n |k|≤n 0
Z 1 X
= f (y) e2πik(x−y) dy ,
0 |k|≤n
| {z }
Dn (x−y)
where
X X
Dn (x) = ek (x) = e2πikx .
|k|≤n |k|≤n
n
X
Dn (x)eπix = eπi(2k+1)x ,
k=−n
n
X n−1
X
Dn (x)e−πix = eπi(2k−1)x = eπi(2k+1)x
k=−n k=−n−1
and hence
so
eπi(2n+1)x − e−πi(2n+1)x sin π(2n + 1)x
Dn (x) = πix −πix
= .
e −e sin πx
Now
Z 1
sn (f, x) = f (y)Dn (x − y) dy
0
Z 1−x
(y−x=t)
= f (x + t)Dn (−t) dt
−x
Z 1−x
(Dn (−t) = Dn (t))
= f (x + t)Dn (t) dt
−x
t7→f (x+t)Dn (t) Z 1/2
has period 1
= f (x + t)Dn (t) dt .
−1/2
We proved
Proposition 6.2. If f ∈ L1 (S 1 ), then
Z 1/2
sn (f, x) = f (x + y)Dn (y) dy ,
−1/2
because
∞ ∞
n−2m+1
Z
X 1 dx
≤ = .
k 2m n x2m 2m − 1
k=n+1
Accordingly
(6.3) ksn − sn0 k∞ ≤ C(m)n−m+1/2 kf (m) k .
This implies the uniform convergence of the sequence of partial sums sn .
Since sn → f in L2 and f is continuous, sn ⇒ f uniformly. Now letting
n0 → ∞ in (6.3) yields
9More precisely, if f ∈ L2 (S 1 ) and |n|m fˆ(n) → 0 as |n| → ∞ for any m, then there is
f ∈ C ∞ (S 1 ) such that f = f˜ a.e.
˜
52 PIOTR HAJLASZ
as |n| → ∞. To prove the opposite implication observe that for any integer
m ≥ 0, |n|m+2 fˆ(n) → 0 as |n| → 0 and hence |nm fˆ(n)| ≤ C(m)|n|−2 for all
n. Thus the series of term by term derivatives
∞
X ∞
X
fˆ(n)e(m)
n = (2πi)m nm fˆ(n)e2πinx
n=−∞ n=−∞
converges uniformly by the M -test. That, however, implies that the Fourier
series of f defines a C ∞ function
∞
X
(6.4) f˜ = fˆ(n)en
n=−∞
with
∞
X
f˜(m) = fˆ(n)e(m)
n
n=−∞
Proof. We have
Z 1/2
s0 + s1 + . . . + sn−1 D0 (y) + . . . + Dn−1 (y)
= f (x + y) dy .
n −1/2 | n
{z }
Fn (y)
and hence
Z 1/2
s0 + s1 + . . . + sn−1
−f = (f (x + y) − f (x))Fn (y) dy .
n −1/2
FUNCTIONAL ANALYSIS 53
where fy (x) = f (x + y). For every ε > 0 we can find 0 < δ < 1/2 such that
I1 (x) < ε/2 for all x by uniform continuity of f . Now
Z Z 1/2
4 sin nπy 2
I2 (x) = ≤ kf k∞ dy
|y|≥δ n δ sin πy
2
≤ (sin πδ)−2 kf k∞ ,
n
and to given δ we can find n0 so large that I2 (x) < ε/2 for n > n0 and all
x ∈ [0, 1]. The proof is complete 2
Fourier coefficients can be defined for f ∈ L1 (S 1 ) by the same integral
formula. Clearly
∧
: L1 (S 1 ) → `∞ (Z)
because
Z 1
|fˆ(n)| = f (x)e−2πinx dx ≤ kf k1 .
0
54 PIOTR HAJLASZ
Thus
s + s + . . . + s
Z 1/2
0 1 n−1
− f
≤ kfy − f k1 Fn (y) dy := I .
n
1 −1/2
For ε > 0 we choose δ > 0 such that kfy − f k1 < ε/2 whenever |y| < δ.
Hence for |y| < δ we can estimate I1 by
Z
ε ε
I1 ≤ Fn (y) dy < .
2 |y|<δ 2
For the second integral note that kfy − f k1 ≤ 2kf k1 and hence
Z
I2 ≤ 2kf k1 Fn (y) dy
|y|≥δ
Z 1/2
1 sin nπy 2
= 4kf k1 dy
δ n sin πy
2
≤ kf k1 (sin πδ)−2
n
which is less than ε/2 provided n is sufficiently large. Now it is easy to see
that the convergence (6.5) implies that the mapping ∧ : L1 (S 1 ) → `∞ (Z) is
one-to-one. The proof is complete. 2
Recall that c0 is a closed subspace of `∞ (Z) consisting of sequences con-
vergent to 0 at ±∞.
Theorem 6.8 (The Riemann-Lebesgue lemma).
∧
: L1 (S 1 ) → c0 ,
i.e. |fˆ(n)| → 0 as |n| → ∞.
FUNCTIONAL ANALYSIS 55
as |n| → ∞. 2
Second proof. If f ∈ C 1 (S 1 ), then
(f 0 )∧ = 2πinfˆ(n)
and hence
fˆ(n) = (2πin)−1 (f 0 )∧ (n) .
Since k(f 0 )∧ k∞ < ∞ it follows that |fˆ(n)| → 0 as |n| → ∞, i.e.
∧
: C 1 (S 1 ) → c0 .
Since10 kfˆk∞ ≤ kf k1 the claim follows from the density of C 1 (S 1 ) in L1 (S 1 ).
2
As mentioned above (cf. Section 9.4), in general, for f ∈ C(S 1 ) the Fourier
series need not converge to f , however, the following results provides suffi-
cient conditions for the convergence.
Theorem 6.9 (Dini’s criterion). If f ∈ L1 (S 1 ) and for some x with |x| ≤
1/2
Z 1/2
f (x + y) − f (x)
dy < ∞ ,
y
−1/2
then
lim sn (f, x) = f (x) .
n→∞
sn (f ) − sn (g) = sn (f − g) ⇒ 0
vanishes for |y| ≤ δ/2 and hence has no singularity at y = 0. Therefore the
function is integrable on S 1 and for |x| ≤ δ/2 we have
Z 1/2
sin(2n + 1)πy
sn (f, x) = f (x + y) dy
−1/2 sin πy
f (x + y) eπiy e2nπiy − e−πiy e−2nπiy
1/2
Z
= dy
−1/2 sin πy 2i
1
(Q+ )∧ (−n) − (Q− )∧ (n) ,
=
2i
where
f (x + y) ±πiy
Q± (y) = e .
sin πy
Proof. We have
Z 1Z 1
∧
(f ∗ g) (n) = f (x − y)g(y) dy e−2πinx dx
0 0
Z 1Z 1
= f (x − y)e−2πin(x−y) dy g(y)e−2πiny dy
0 0
= fˆ(n)ĝ(n)
The proof is complete. 2
Now we will show several applications of the Furier series.
The equality is achieved for the function f (x) = 2i sin 2πx = e2πix − e−2πix .
2
Proof. We can assume that the fixed length of the Jordan curve is 1.
Denoting the enclosed area by A we can write the theorem in the form of
the inequality
1
(6.7) A≤
4π
with the equality if and only if the curve is a circle. We will assume that the
Jordan curve γ(t) = (x(t), y(t)) is of class C 1 (S 1 ), i.e. γ is
closed:
x(0) = x(1), y(0) = y(1);
smooth:
x, y ∈ C 1 (S 1 ), γ(t) 6= 0 for all t;
Jordan:
γ(t1 ) 6= γ(t2 ) for t1 6= t2 , t1 , t2 ∈ [0, 1);
unit length:
Z 1 1/2
`= ẋ(t)2 + ẏ(t)2 dt = 1 .
0
60 PIOTR HAJLASZ
Therefore
(x(t) − x̂(0))2 + (y(t) − ŷ(0))2 = const.
The proof is complete. 2
and
n−1 n−1
1X 1 X 2πim(x+kγ)
em (xk ) = e
n n
k=0 k=0
1 e2πimx (1 − e2πimγn )
( geom.
sum
)
= → 0 as n → ∞ .
n 1 − e2πimγ
(Note that sice γ is irrational the denominator is nonzero.) Thus (6.11) is
satisfied. Next (6.11) is satisfied by trigonometric polynomials which are
finite sums of the form
X
(6.12) f (x) = am em (x)
|m|≤k
1 n−1 1
Z
X
+ (fε (xk ) − f (xk )) + (fε − f )
n
| k=0 {z } | 0 {z }
<ε/3
<ε/3
1 n−1 Z 1
X 2ε
≤ fε (xk ) − fε + < ε.
n 0 3
k=0
The proof is complete. 2
Corollary 6.18. If 0 < γ < 1 is irrational and 0 ≤ a < b ≤ 1, then
#{k < n : a ≤ xk ≤ b}
lim = b − a.
n→∞ n
(a) Investigate how many times 7 and 8 will appear as a first digit of the
decimal representation of 2n for n ≤ 45.
(b) The digit c appears as a first digit of 2n with the frequency
#{k < n : 2k = c . . .}
lim .
n→∞ n
Prove that this limit exists for any c ∈ {1, 2, . . . , 9} and show that 7
appears more often than 8.
7. Spherical harmonics
mutually orthogonal.
It suffices to prove that the orthogonal complement of the subspace |x|2 Pk−2
is Hk , i.e.
(7.3) Pk = Hk ⊕ |x|2 Pk−2
because it readily implies unique representation (7.1). To prove (7.3) observe
that P1 ∈ Pk is in the orthogonal complement of |x|2 Pk−2 if and only if for
every P2 ∈ Pk−2
h|x|2 P2 , P1 i = 0 .
We have
∂ ∂
0 = h|x|2 P2 , P1 i = ∆ P2 P1 = P2 ∆P1 = hP2 , ∆P1 i
∂x ∂x
for every P2 ∈ Pk−2 , where in the right hand side we have the inner product
in Pk−2 . This is, however, true if and only if ∆P1 = 0, i.e. P1 ∈ Hk . 2
Corollary 7.3. The restriction of any polynomial to the unit sphere S n−1
is a finite sum of surface spherical harmonics.
We will prove now that the surface spherical harmonics are eigenfunctions
of the spherical Laplacean ∆S .
Theorem 7.5. If Y ∈ Hk , then
∆S Y (x) = −k(k + n − 2)Y (x) .
Since Pk = |x|k Ỹk ∈ Hk is harmonic, the estimate will follow from suitable
estimates for harmonic functions.
Recall that harmonic functions have the mean value property, i.e. if u is
harmonic on Rn , then
Z
u(x) = u dσ
S n−1 (x,r)
for every x ∈ Rn and every r > 0. This easily implies Rthat if ϕ ∈
C0∞ (B n (0, 1)) is radial (i.e. constant on spheres centered at 0), Rn ϕ(x) dx =
1 and ϕε (x) = ε−n ϕ(x/ε), then
u(x) = (ϕε ∗ u)(x)
for every ε > 0 and every x ∈ Rn . Thus
Dα u(x) = (Dα ϕε ∗ u)(x)
Now
Z Z Z 1+ε
2 2
|Pk (x)| dx = |Yk | dσ t2k+n−1 dt ≤ (1+ε)2k+n kYk k2L2 .
|x|≤1+ε S n−1 0
Since Ỹk (x) = |x|−k Pk (x), the Leibnitz rule implies (7.6). 2
The Baire category theorem proved below plays an important role in many
areas of mathematics. In this section we will show its applications outside
functional analysis and in the next two sections we will use it in the proofs
of two fundamental theorems in functional analysis, the Banach-Stenihaus
theorem and the Banach open mapping theorem.
Theorem 8.1 (Baire). The intersection of a countable family of open and
dense sets in a complete metric space is a dense set.
Since [a, b] is complete, it follows from the Baire theorem that for some n
the set En ∩ [a, b] has nonempty interior (in the topology of [a, b]), so there is
(c, d) ⊂ En ∩[a, b] such that f (n) = 0 on (c, d). Accordingly f is a polynomial
on (c, d) and hence
(c, d) ⊂ Ω ∩ [a, b] 6= ∅.
The set X = R \ Ω is closed and hence complete. It remains to prove that
X = ∅. Suppose not. Observe that every point x ∈ X is an accumulation
point of the set, i.e. there is a sequence xi ∈ X, xi 6= x, xi → x. Indeed,
otherwise x would be an isolated point, i.e. there would be two intervals
(8.3) (a, x), (x, b) ⊂ Ω, x 6∈ Ω .
The function f restricted to each of the two intervals is a polynomial, say of
degrees n1 and n2 . If n > max{n1 , n2 }, then f (n) = 0 on (a, x) ∪ (x, b). Since
f (n) is continuous on (a, b), it must be zero on the entire interval and hence
f is a polynomal of degree ≤ n − 1 on (a, b), so (a, b) ⊂ Ω which contradicts
(8.3).
The space X = R \ Ω is complete. Since
∞
[
X= X ∩ En ,
n=1
We will prove that f (n) = 0 on (a, b). This will imply that (a, b) ⊂ Ω which
is a contradiction with (8.4). Since f (n) = 0 on X∩(a, b) = (a, b)\Ω it remains
to prove that f (n) = 0 on (a, b)∩Ω. To this end it suffices to prove that for any
interval (ai , bi ) that appears in (8.2) such that (ai , bi )∩(a, b) 6= ∅, f (n) = 0 on
(ai , bi ). Since (a, b) is not contained in (ai , bi ) one of the endpoints belongs
to (a, b), say ai ∈ (a, b). Clearly ai ∈ X ∩ (a, b) and hence f (m) (ai ) = 0 for
all m ≥ n. If f is a polynomial of degree k on (ai , bi ), then f (k) is a nonzero
constant on (ai , bi ), so f (k) (ai ) 6= 0 by continuity of the derivative. Thus
k < n and hence f (n) = 0 on (ai , bi ). 2
Exercise. As the previous exercise shows the theorem is not true if we
only assume that f ∈ C 1000 . Where did we use in the proof the assumption
f ∈ C ∞ (R)?
FUNCTIONAL ANALYSIS 73
Theorem 8.7 (Banach). The set of function in C[0, 1] that are not differ-
entiable at any point of [0, 1] is a dense Gδ subset of C[0, 1]. In particular it
is not empty.
Proof of the theorem. Let f ∈ C[0, 1], x ∈ [0, 1] and 0 < r ≤ 1/2. If both
numbers x + r, x − r belong to the interval [0, 1], then we define
n |f (x + r) − f (x)| |f (x − r) − f (x)| o
(8.5) D(f, x, r) = min , .
r r
T∞
It is easy to see that if f ∈ n=1 Gn , then
f (x + r) − f (x)
lim sup
= +∞ for all x ∈ [0, 1]
r→0 r
and hence f is not differentiable at any point of [0, 1]. Thus it remains to
prove that everyTset Gn is open and dense, because then the Baire theorem
will imply that ∞ 21
n=1 Gn is a dense Gδ subset of C[0, 1]. The fact that the
sets Gn are open is easy and left to the reader. To prove density it suffices
to prove that every g ∈ C[0, 1] can be uniformly approximated by functions
g̃ ∈ C[0, 1] such that
Then we modify it to a piecewise linear function which has flat parts and
parts with the slope larger than n + 1.
Finally we add little teeth to the flat part, so the slope of each tooth is
also larger than n + 1.
The proof is based on the following general idea. Given a complete metric
space X, we want to prove that there is an element x ∈ X that has a certain
property P . We find other, simpler to deal with, properties Pn , n = 1, 2, . . .
such that
x∈ ∞
T
Then
T∞ n=1 Gn has the property P and such an x exists, because the
set n=1 Gn is nonempty by the Baire theorem. This is what is called the
Baire category method.
Exercise. We say that a function f ∈ C ∞ (0, 1) is analytic at a ∈ (0, 1) if
there is ε > 0 such that f (x) = ∞ (n) (a)(x − a)n /n! for |x − a| < ε. Use
P
n=0 f
the Baire category method to prove that there is a function f ∈ C ∞ (0, 1)
that is not analytic at any point.
9. Banach-Steinhaus theorem
The following theorem as well as each of the four corollaries that follow
are called Banach-Steinhaus theorem.
Theorem 9.1 (Banach-Steinhaus). Let X be a complete metric space and
let {fi }i∈I be a family of continuous real-valued functions on X. Then exactly
one of the following two conditions is satisfied
then there is a nonempty open set U ⊂ X and a constant M > 0 such taht
sup |fi (x)| ≤ M for all x ∈ U .
i∈I
The above two results are very important in the case X is a Banach space.
Corollary 9.3. Let X be a Banach space and Y a normed space. Let
{Li }i∈I ⊂ B(X, Y ). Then either
sup kLi k < ∞
i∈I
or there is a dense Gδ set E ⊂ X such that
sup kLi xk = ∞ for all x ∈ E .
i∈I
Proof. The functions fi (x) = kLi xk, i ∈ I are continuous and real-valued.
If the second condition is not satisfied, then it follows from the Banach-
Steinhaus theorem that there is an open ball B(x0 , r0 ) ⊂ X such that
sup |fi (x)| = M < ∞ for all x ∈ B(x0 , r0 ) .
i∈I
For x 6= 0
r0
y = x0 + x ∈ B(x0 , r0 )
2kxk
and hence for all i ∈ I we have
2kxk
r0
4M
kLi xk =
Li x0 + x − Li x0
≤ kxk
r0 2kxk r0
which yields
4M
sup kLi k ≤ < ∞.
i∈I r0
The claim is proved. 2
Corollary 9.4. Let X be a Banach space and Y a normed space. Let
{Li }i∈I ⊂ B(X, Y ). If for every x ∈ X
sup kLi xk < ∞
i∈I
then
sup kLi k < ∞ .
i∈I
and
kLk ≤ lim inf kLn k .
n→∞
and thus
sup kTn k = M < ∞
n
by the Banach-Steihnaus theorem. Hence |Tn (y)| ≤ M kyk for every n and
all y ∈ Y . In particular
Proof. We will prove the theorem in the case 1 < p ≤ ∞, but an obvi-
ous modification gives also the proof in the case p = 1. Define a bounded
functional on `p by the formula
n
X
Tn x = ηi ξi , n = 1, 2, 3, . . . ,
i=1
For example if
1
0 0 0 ...
1 0 1
1
1 1
0 0 ...
1
2 2
A1 = A2 =
1
1 1 1
0 ...
. .
3 3 3
.
0 ...
then the sequence (an ) is summable to g by the method A1 if an → g and
by the method A2 if
a1 + a2 + . . . + an
→ g,
n
so we can recover both the classical notion of the convergence and the Cesaro
summability method.
Definition. We say that the matrix method A is regular if every convergent
sequence if summable to the same limit by the method A.
Clearly methods A1 and A2 are regular.
Theorem 9.9 (Toeplitz). The matrix summability method A = (ξij )∞
i,j=1 is
regular if and only if the following conditions are satisfied:
(a) supn ∞
P
i=1 |ξni | < ∞,
(b) limn→∞ P ξni = 0 for i = 1, 2, 3, . . .,
(c) limn→∞ ∞ i=1 ξni = 1.
FUNCTIONAL ANALYSIS 81
It remains to prove that conditions (a), (b), (c) imply that the method A
is regular.
For every x = (an ) ∈ c the series
∞
X
Tn x = ξni ai
i=1
Theorem 9.10. There is a dense Gδ set E ⊂ C(S 1 ) such that for each
f ∈ E the set
{x ∈ [0, 1] : sup |sn (f, x)| = +∞}
n
where
sin π(2n + 1)y
Dn (y) = .
sin πy
Clearly
Z 1/2
(9.4) kΛn k ≤ |Dn (y)| dy .
−1/2
Indeed, let
1 if Dn (y) ≥ 0,
g(y) =
−1 if Dn (y) < 0.
In the above reasoning we could replace 0 by any other point x ∈ [0, 1], so
for each x ∈ [0, 1] there is a dense Gδ set Ex ⊂ C(S 1 ) such that
sup |sn (f, x)| = ∞ for all f ∈ Ex .
n
Accordingly supn |sn (f, x)| = ∞ for x in a dense Gδ set, which is uncount-
able by Theorem 8.4. 2
N → X/M n 7→ [n]
are one-to-one and onto, hence they are isomorphisms of Banach spaces. 2
Proposition 10.7. If L : X ⊕ Y → Z is an isomorphism of Banach spaces,
then L(X) and L(Y ) are closed subspaces of Z and hence Z = L(X)⊕L(Y ).
88 PIOTR HAJLASZ
Proof. H = M ⊕ M ⊥ . 2
Definition. Let X be a Banach space. The mapping P ∈ B(X) = B(X, X)
is called a projection if P 2 = P , i.e. P (P (x)) = P (x) for all x ∈ X. We denote
the kernel (null space) and the range of the projection by
N (P ) = {x ∈ X : P x = 0}, R(P ) = {P x : x ∈ X} .
Theorem 10.9.
This is a very difficult theorem and we will not prove it. We will prove,
however
Theorem 10.13 (Phillips). The space c0 is not complemented in `∞ . Equiv-
alently there is no bounded linear projection of `∞ onto c0 .
Proof. It suffices to prove the lemma with N replaced by Q ∩ (0, 1). For
each irrational number i ∈ (0, 1) \ Q := I let Ai be a sequence of rationals
in (0, 1) convergent to i. It is easy to see that the family {Ai }i∈I has the
desired properties. 2
By contradiction suppose that
`∞ = c0 ⊕ X, X ⊂ `∞ ,
so X ' `∞ /c0 . Consider the family of functionals {en }n∈N on X
hen , xi = xn , x = (xi ) ∈ X ⊂ `∞ .
The family {en }n in X ∗ is total, i.e. it has the property that
hen , xi = 0 for all n ⇒ x = 0.
It suffices to prove that there is no countable total family of functionals in
(`∞ /c0 )∗ .
90 PIOTR HAJLASZ
We will prove now that for every x∗ ∈ (`∞ /c0 )∗ the set
{[fi ] : hx∗ , [fi ]i =
6 0}
is countable. To this end it suffices to prove that for each integer n the set
C(n) = {[fi ] : hx∗ , [fi ]i ≥ 1/n}
is finite. Choose [fi1 ], . . . , [fim ] ∈ C(n) and let
is countable and hence there is24 fi such that hhn , [fi ]i = 0 for all n, so the
family {hn }n cannot be total. 2
Let
α = sup f˜(x) − p(x − x1 ) .
x∈M̃
Then
(11.1) f˜(x) − α ≤ p(x − x1 ) and f˜(y) + α ≤ p(x1 + y) .
˜ by the formula
We extend f˜ to M̃
˜
f˜(x + tx1 ) = f˜(x) + tα .
Replacing x and y by x/(−t), t < 0 and x/t, t > 0 in (11.1), after simple
calculations we arrive at
˜ ˜.
f˜(x) ≤ p(x) on M̃
Now we want to apply the Hausdorff maximality theorem. Consider the
family of pairs (M̃ , f˜), where M̃ ⊃ M is a subspace of X and f˜ : M̃ → R is
an extension of f satisfying
f˜(x) ≤ p(x) for x ∈ M̃ .
The family is partially ordered by the relation
˜ f˜˜)
(M̃ , f˜) ≤ (M̃,
˜ and f˜˜ is an extension of f˜. According to the Hausdorff maximality
if M̃ ⊂ M̃
theorem there is a maximal element in the family. Because of the first part
of the proof the maximal element must be a functional defined on all of X.
Denote it by (X, F ). Hence
F (x) = f (x) for x ∈ M , F (x) ≤ p(x) for x ∈ X.
Now it suffices to observe that
F (x) = −F (−x) ≥ −p(−x)
by linearity of F . 2
Theorem 11.2 (Hahn-Banach). Let M be a subspace of a linear space X
over K (= R of C), and let p be a seminorm25 on X. Let f be a linear
functional on M such that
|f (x)| ≤ p(x) for x ∈ M .
Then there is an extension F : X → K of f such that
F (x) = f (x) for x ∈ M ; |F (x)| ≤ p(x) for x ∈ X.
25i.e. p(x + y) ≤ p(x) + p(y), p(αx) = |α|p(x) for x, y ∈ X and α ∈ K, see Section 4.2.
FUNCTIONAL ANALYSIS 93
27Rn or Cn .
FUNCTIONAL ANALYSIS 97
(b)
LIM (ax(t) + by(t)) = a LIM x(t) + b LIM y(t) ;
t→∞ t→∞ t→∞
(c)
LIM x(t + τ ) = LIM x(t) for every τ > 0 ;
t→∞ t→∞
(d)
lim inf x(t) ≤ LIM x(t) ≤ lim sup x(t) .
t→∞ t→∞ t→∞
and set
LIM xn := LIM x(t) .
n→∞ t→∞
and then
p(x) = inf β(x; t1 , . . . , tn ) : ti ≥ 0, n ∈ N .
We will prove that p(x) is a Banach functional. Clearly p(tx) = tp(x) for
t > 0 and we only need to prove subadditivity p(x + y) ≤ p(x) + p(y).
Let t1 , . . . , tn ≥ 0 and s1 , . . . , sm ≥ 0. Then ti + sj is a collection of nm
numbers. Denote these numbers by u1 , . . . , unm . It it easy to see that the
subadditivity of p follows from the lemma.
Lemma 11.11.
β(x + y; u1 , . . . , unm ) ≤ β(x; t1 , . . . , tn ) + β(y; s1 , . . . , sm ) .
FUNCTIONAL ANALYSIS 99
Proof. We have
nm m n n m
1 X 1 X1X 1X 1 X
(x+y)(t+uk ) = x(t+ti +sj )+ y(t+ti +sj ) .
nm m n n m
k=1 j=1 i=1 i=1 j=1
= β(x; t1 , . . . , tn ) + β(y; s1 , . . . , sm ) .
This completes the proof of the lemma and hence that of the fact that p(x)
is a Banach functional. 2
If x ∈ M ⊂ B(0, ∞), then
lim x(t) = p(x) ≤ p(x)
t→∞
and hence according to the Hahn-Banach theorem the functional x 7→
limt→∞ x(t) extends from M to a linear functional on B [0, ∞) that we
denote by LIM t→∞ x(t) such that
−p(−x) ≤ LIM x(t) ≤ p(x) for all x ∈ B [0, ∞).
t→∞
11.2. Finitely additive measures. In this section we will prove the fol-
lowing surprising result.
Theorem 11.12 (Banach). There is a finitely additive measure µ : 2R →
[0, ∞], i.e.
µ(A ∪ B) = µ(A) + µ(B) for A, B ⊂ R, A ∩ B = ∅
defined on all subsets of R and such that
Banach proved that the above result holds also in R2 , i.e. there is a finitely
additive measure
2
µ : 2R → [0, ∞]
defined on all subsets of R2 , invariant under isometries of R2 and equal to
the Lebesgue measure on the class of Lebesgue measurable sets. According
to the Vitali example such measures cannot be, however, countably additive.
Surprisingly, the Banach theorem does not hold in Rn , n ≥ 3. This is
related to an algebraic fact that the group of isometries of Rn , n ≥ 3 contains
a free supgroup of rank 2. Namely Banach and Tarski proved30 that the unit
ball in R3 can be decomposed into a finite number of disjoint sets (later it
was shown that it suffices to take 5 sets)
B 3 (0, 1) = A1 ∪ A2 ∪ A3 ∪ A4 ∪ A5
is a way that there are isometries τ1 , . . . , τ5 of R3 such that
B 3 (0, 1) = τ1 (A1 ) ∪ τ2 (A2 ), B 3 (0, 1) = τ3 (A3 ) ∪ τ4 (A4 ) ∪ τ5 (A5 )
and a similar decomposition is possible for B n (0, 1) for any n ≥ 3.
Clearly the sets Ai cannot be Lebesgue measurable, because we would
obtain a contradiction by comparing volumes. This also shows that there
is no finitely additive measure in Rn , n ≥ 3 invariant under isometries and
equal to the Lebesgue measure on the class of Lebesgue measurable sets.
The Banach theorem will follow from a somewhat stronger result. As in
the case of Fourier series we can identify bounded functions on R with period
1 with bounded functions on S 1 via the exponential mapping t R7→ e2πit . Thus
1
the integral of a function on S 1 corresponds to the integral 0 x(t) dt of a
function x : R → R with period 1.
30See S. Wagon, The Banach-Tarski paradox, Cambridge Univ. Press 1999.
FUNCTIONAL ANALYSIS 101
(a) Z Z Z
ax(t) + by(t) dt = a x(t) dt + b y(t) dt ;
(b) Z
x(t) dt ≥ 0 if x ≥ 0
(c) Z Z
x(t + τ ) dt = x(t) dt for all τ ∈ R;
(d) Z Z
x(1 − t) dt = x(t) dt ;
As in the proof of Theorem 11.10 one can show that p(x) is a Banach func-
tional such that
(11.7) p(xτ − x) ≤ 0, p(x − xτ ) ≤ 0 .
For a Lebesgue measurable function x ∈ B(S 1 ) we define a functional
Z 1
∗
hx1 , xi = x(t) dt .
0
It is easy to see that
hx∗1 , xi ≤ p(x)
and hence x∗1 can be extended to a functional x∗ on B(S 1 ) such that
−p(−x) ≤ hx∗ , xi ≤ p(x) for all x ∈ B(S 1 ).
Finally we define Z
1
x(t) dt = hx∗ , x + x− i ,
2
where x− (t) = x(1−t). The properties (a), (d) and (e) are obvious. Property
(b) follows from the observation that if x ≥ 0, then p(−x) ≤ 0 and hence
0 ≤ −p(−x) ≤ hx∗ , xi, 0 ≤ −p(−x− ) ≤ hx∗ , x− i .
Finally, property (c) follows from the inequality (11.7). 2
and the corresponding integral for any complex polynomial equals 0. Note
that the domain Ω is not simply connected. Hence there is no complex
version of the Weierstrass approximation theorem. However, we have
Theorem 11.14 (Runge). If Ω ⊂ C is simply connected, then every holo-
morphic function on Ω can be uniformly approximated on compact subsets
of Ω by complex polynomials.
The integral on the right hand side can be uniformly approximated by Rie-
mann sums
Z m
1 f (ξ) 1 X f (ξk )
dξ ≈ ∆ξk
2πi γ ξ − z 2πi ξk − z
k=1
which is a linear combination of functions
1
z 7→ ∈ C(K) .
ξk − z
Thus it remains to prove that every function
1
(11.8) z 7→ ∈ C(K), ξ0 6∈ K
ξ0 − z
can be uniformly approximated on K by complex polynomials. Let M ⊂
C(K) be the closure of the subspace of complex polynomials. Suppose that
the function (11.8) does not belong to M . According to Theorem 11.6 there
is a functional x∗ ∈ C(K)∗ which vanishes on M and is nonzero on the
function (11.8), i.e.
hx∗ , p(z)i = 0, p any complex polynomial,
1
(11.9) hx∗ , i=
6 0.
ξ0 − z
The function
1
(11.10) ξ 7→ hx∗ , i
ξ−z
is complex differentiable in C \ K, so it is holomorphic in C \ K. If |ξ| > R =
maxz∈K |z|, then
∞
1 1 1 X zn
= =
ξ−z ξ 1 − z/ξ ξ n+1
n=0
and hence the function
1
z 7→
ξ−z
can be uniformly approximated in C(K) by complex polynomials, so
1
hx∗ , i=0 if |ξ| > R.
ξ−z
Since the function (11.10) is homomorphic in C \ K, C \ K is connected and
it vanishes for ξ sufficiently large
1
hx∗ , i=0 for all ξ ∈ C \ K
ξ−z
which contradicts (11.9). 2
104 PIOTR HAJLASZ
In Theorem 11.17 we will show that the conditions from the proposition
characterize the unit ball for a seminorm.
Definition. For each convex and absorbing set W ⊂ X the Minkowski
functional is n x o
µW (x) = inf s > 0 : ∈W .
s
Theorem 11.16. Let W ⊂ X be convex and absorbing. The Minkowski
functional has the following properties
(a) µW (x) ≥ 0;
(b) µW (x + y) ≤ µW (x) + µW (y);
(c) µW (tx) = tµW (x) for all t ≥ 0;
(d) µW (x) = 0 if and only if {tx : t ≥ 0} ⊂ W .
(e) If x ∈ W , then µW (x) ≤ 1;
(f) If µW (x) < 1, then x ∈ W
(g) If µW (x) > 1, then x 6∈ W .
Proof. All the properties but (b) are obvious. To prove (b) let x, y ∈ X
and s1 , s2 > 0 be such that x/s1 , y/s2 ∈ W . It follows from the convexity of
W that
x+y s1 x s2 y
= + ∈W
s1 + s2 s1 + s2 s1 s1 + s2 s2
Hence µW (x + y) ≤ s1 + s2 and the claim follows upon taking the infimum.
2
31As always K = R or K = C.
32Equivalently a convex set W ⊂ X is absorbing if X = S∞ nW .
n=1
FUNCTIONAL ANALYSIS 105
Note that the properties (b) and (c) show that the Minkowski functional
is a Banach functional.
The following result provides a geometric characterization of sets that are
unit balls for norms and seminorms.
Theorem 11.17. If W ⊂ X is convex, balanced and absorbing at any point
of W , then there is a unique seminorm p such that W = {x ∈ W : p(x) < 1}.
Moreover p is a norm if and only if for each x 6= 0 there is t > 0 such that
tx 6∈ W .
linear space, the real part of a functional is R-linear and we can still apply
the Hahn-Banach theorem (cf. Theorem 11.2).
Theorem 11.18. Let W1 , W2 be disjoint convex subsets of a normed space
X and assume that W1 is open. Then there is a functional x∗ ∈ X ∗ and
c ∈ R such that
re hx∗ , xi < c ≤ re hx∗ , yi
for all x ∈ W1 and y ∈ W2 .
Proof. It suffices to prove the theorem in the case of real normed spaces.
Indeed, if X is a complex normed space, then it can still be regarded as a
real normed space. If we can find a bounded R-linear functional x̃∗ such that
hx̃∗ , xi < c ≤ hx̃∗ , yi
for all x ∈ W1 and y ∈ W2 , then
hx∗ , xi = hx̃∗ , xi − ihx̃∗ , ixi
is a bounded C-linear functional that satisfies re hx∗ , xi = hx̃∗ , xi.
Henceforth we will assume that X is a real normed space. Fix x1 ∈ W1
and x2 ∈ W2 . Then the set
W = W1 − W2 + x2 − x1
| {z }
x0
so
(11.13) hx∗ , xi < hx∗ , yi for all x ∈ W1 , y ∈ W2 .
The sets x∗ (W1 ) and x∗ (W2 ) are intervals in R as convex sets. Moreover
x∗ (W1 ) is open, since W1 is open.33 Therefore (11.13) yields the existence of
c ∈ R such that x∗ (W1 ) ⊂ (−∞, c), x∗ (W2 ) ⊂ [c, ∞). 2
Theorem 11.19. Let W1 , W2 be disjoint convex subsets of a normed space
X. Assume that W1 is compact and W2 is closed. Then there is x∗ ∈ X ∗ ,
c ∈ R and ε > 0 such that
re hx∗ , xi ≤ c − ε < c ≤ re hx∗ , yi
for all x ∈ W1 and y ∈ W2 .
34A metric space is totally bounded if for every ε > 0 there is a finite covering of the
space by balls of radius ε.
FUNCTIONAL ANALYSIS 109
and let W = L(f (E)). Since L is linear the set L(K) is compact, convex
and it contains W , so co(W ) ⊂ L(K). It suffices to prove that
Z Z
z = (z1 , . . . , zn ) := hx∗1 , f i dµ, . . . , hx∗n , f i dµ ∈ co(W ) .
E E
110 PIOTR HAJLASZ
Indeed, since co(W ) ⊂ L(K) it will imply that z = L(y) for some y ∈ K
which is (12.4).
Suppose z 6∈ co(W ). The sets {z} and co(W ) are convex, compact and
disjoint, so according to Theorem 11.19 there is a functional in (Kn )∗ that
separates the two sets, i.e. there is (c1 , . . . , cn ) ∈ Kn such that
n
X n
X
re ci ti < re ci zi for all t = (t1 , . . . , tn ) ∈ co(W ).
i=1 i=1
If t = L(f (s)), s ∈ E, we have
n
X n
X
re ci hx∗i , f (s)i < re ci zi .
i=1 i=1
Since µ is a probability measure, integration of the inequality over E gives
X n Z Xn
re ci hx∗i , f i dµ < re ci zi
i=1 E i=1
| {z }
zi
which is an obvious contradiction. The proof of the theorem is complete. 2
Theorem 12.3. Under the assumptions of Theorem 12.1
Z
Z
f dµ
≤ kf k dµ .
E E
R
Proof. Let y = E f dµ. According to the Hahn-Banach theorem there is
x∗ ∈ X ∗ such that kx∗ k = 1 and hx∗ , yi = kyk. Hence
Z
Z Z
f dµ
= hx∗ , yi = hx∗ , f i dµ ≤ kf k dµ .
E E E
The proof is complete. 2
by
hx∗∗ ∗ ∗∗ ∗ −1 ∗
0 , x i = hy0 , (T ) x i .
Since X is reflexive, there is x0 ∈ X such that
hx∗∗ ∗ ∗ ∗
0 , x i = hx , x0 i for x ∈ X
∗
and hence
hy0∗∗ , y ∗ i = hy0∗∗ , (T ∗ )−1 (T ∗ y ∗ )i = hx∗∗ ∗ ∗
0 ,T y i
= hT ∗ y ∗ , x0 i = hy ∗ , T x0 i .
Thus (13.3) holds with y0 = T x0 ∈ Y . 2
Corollary 13.5. Finitely dimensional Banach spaces are reflexive.
Theorem 13.6. X is reflexive if and only if X ∗ is reflexive.
i.e. xni = δni , then kxn − xm kp = 21/p and hence no subsequence of xn can
be convergent in the norm. However, xn is weakly convergence to 0. Indeed,
according to Theorem 2.12 for every x∗ ∈ (`p )∗ there is s = (si ) ∈ `q such
that
∞
X
hx∗ , xn i = si xni = sn
i=1
(a)
|fn (x)| ≤ M for all x ∈ X and n = 1, 2, 3 . . .
(b)
fn (x) → f (x) for all x ∈ X.
Suppose x0 6∈ Wn . The set Wn is convex and closed and the set W 0 = {x0 }
is convex and compact. According to Theorem 11.19 there is x∗ ∈ X ∗ , c ∈ R
and ε > 0 such that
re hx∗ , x0 i ≤ c − ε < c ≤ re hx∗ , xi
for all x ∈ Wn . In particular
re hx∗ , x0 i ≤ c − ε < c ≤ re hx∗ , xi i for i = n, n + 1, . . .
which contradicts weak convergence xi * x0 . 2
Definition. Let X be a normed space. A sequence of functionals {x∗n } ⊂ X ∗
converges weakly-∗ to x∗0 ∈ X ∗ if for every x ∈ X
hx∗n , xi → hx∗0 , xi as n → ∞.
∗
We denote weak-∗ convergence by x∗n * x∗0 .
∗
Theorem 14.5. If x∗n * x∗0 , then
sup kx∗n k < ∞ and kx∗0 k ≤ lim inf kx∗n k .
n n→∞
Corollary 14.7. Let X be a locally compact metric space and let {µn }n be
a sequence of finite positive Borel measures on X such that
sup µn (X) < ∞ .
n
Then there is a subsequence µnk and a finite Borel measure µ on X such
that
µ(X) ≤ lim inf µnk (X)
k→∞
and for every f ∈ C0 (X)
Z Z
f dµnk → f dµ .
X X
The following result generalizes Theorem 5.1 to any reflexive space, see
also an example that follows Theorem 5.1.
Theorem 14.10. Every nonempty, convex and closed set E in a reflexive
space X contains an element of smallest norm.
Proof. Let {xn } ⊂ E be a sequence such that kxn k → inf x∈E kxk. Hence
the sequence {xn } is bounded, so it has a weakly convergent subsequence
Hence z0 = x0 − y0 , x0 ∈ X0 satisfies
kx0 − y0 k = inf kx − y0 k
x∈X0
and the proof of the Riesz lemma shows that the vector y = (y0 − x0 )/ky0 −
x0 k satisfies the claim. 2
The following result which is interesting on its own will be useful later.
Proposition 15.1. Let s be be the space of all (real or complex) sequences
x = (xk )∞
k=1 . The space s with the metric
∞
X |xk − yk |
d(x, y) = 2−k
1 + |xk − yk |
k=1
is a complete metric space. Moreover for each sequence of positive numbers,
r1 , r2 , . . . > 0 the set
K = {x ∈ s : |xk | ≤ rk for k = 1, 2, . . .}
is compact in (s, d).
That means for each bounded set E ⊂ X ∗ , the metric ρE being the
restriction of ρ to E is such that the convergence in ρE is equivalent to the
weak-∗ convergence in E, i.e. the weak-∗ convergence in bounded subsets of
X ∗ is metrizable. This is true in bounded sets only as the weak-∗ convergence
in X ∗ , dim X = ∞ is not metrizable (see Theorem 15.4).
The following is a version of the separable Banach-Alaoglu theorem (The-
orem 14.6).
Corollary 15.3. If X is a separable Banach space, then every closed ball
B = {x∗ ∈ X ∗ : kx∗ k ≤ r}
is compact in the metric ρB .
∗
If x∗n * x∗ , then hx∗n , xi i → hx∗ , xi i for every i, so every coordinate
of Φ(x∗n ) converges to the corresponding coordinate of Φ(x∗ ) and hence
d(Φ(x∗n ), Φ(x∗ )) → 0.
The assumption that E is bounded was employed only once in the proof
of the implication from (15.1) to (15.2) and the boundedness was a crucial
assumption here as the following result shows.
Xn = span {x1 , x2 , . . . , xn }
kx − xm k < ε .
(a) ∅, X ∈ T ;
(b) T is closed under finite intersections, i.e. if U, V ∈ T , then US∩V ∈ T ;
(c) T is closed under arbitrary unions, i.e. if {Ui }i∈I ⊂ T , then i∈I Ui ∈
T.
Proposition 15.5.
Proof. We leave the proofs of (a)-(c) as an exercise and we will only prove
property (d). We need to show that for every open set U ⊂ X, (f −1 )−1 (U ) =
f (U ) is an open subset of f (X). X \ U is closed, so it is compact by (a).
Hence f (X \ U ) is compact by (c). Since f (X) is Hausdorff, f (X \ U ) is
closed and hence f (U ) = f (X) \ f (X \ U ) is open. 2
Definition. Let K be a family of functions from a set Y into a topological
space (X, T ). The K-weak topology in Y is the weakest topology on Y for
which all the functions in the family K are continuous.
The K-weak topology is constructed as follows. Observe that the sets
B = {f1−1 (U1 ) ∩ . . . ∩ fn−1 (Un ) : fi ∈ K, Ui ∈ T , i = 1, 2, . . . , n}
must be open and that the family of all possible unions of the sets from B
is a topology, so it is the K-weak topology and B is a base.
Let (X1 , T1 ), . . . , (Xn , Tn ) be topological spaces. In the Cartesian product
n
Y
Xi = X1 × . . . × Xn
i=1
46The claim is not true if the space is not Hausdorff. Indeed, {∅, X} is a topology in
X and every subset is compact, but only ∅ and X are closed.
128 PIOTR HAJLASZ
If (X1 , d1 ), . . . , (Xn , dn ) are metric spaces, thenQit is easy to see that the
topology induced by the metric d in the product ni=1 Xi
n
X
d((x1 , . . . , xn ), (y1 , . . . , yn )) = di (xi , yi )
i=1
is equipped with the product topology which is the weakest topology for
which each of the projection
Y
πj : Xi → Xj , πj ((xi )i∈I ) = xj
i∈I
is continuous.
Let i1 , . . . , in ∈ I and Ui1 ∈ Ti1 , . . . , Uin ∈ Tin be chosen arbitrarily. It is
easy to see that the sets
{(xi )i∈I : xi1 ∈ Ui1 , . . . , xin ∈ Uin } = πi−1
1
(Ui1 ) ∩ . . . ∩ πi−1
n
(Uin )
form a base for the product topology.
Q
It is easy to see that if the spaces {Xi }i∈I are Hausdorff, then i∈I Xi is
Hausdorff.
Theorem 15.6 (Tychonov). Let {Xi }i∈I Q be an arbitrary collection of com-
pact topological spaces. Then the product i∈I Xi is compact.
Proof. Let
Φ : X ∗ → s, Φ(x∗ ) = (hx∗ , x1 i, hx∗ , x2 i, . . .)
be a mapping defined in the proof of Theorem 15.2. The mapping is one-to-
one. It remains to prove that Φ is continuous. Indeed, since s is Hausdorff
as a metric space and B is compact in the weak-∗ topology, it will follow
from Proposition 15.5(d) that
Φ|B : B → s
is a homeomorphism onto the image and hence the weak-∗ topology in
B will coincide with the topology generated by the metric ρ(x∗ , y ∗ ) =
d(Φ(x∗ ), Φ(y ∗ )).
Fix r > 0 and let w = (wi ) ∈ s. Let N be such that ∞ −k < r/2.
P
k=N +1 2
Then the set
A(w, r, N ) = {t ∈ s : |ti − wi | < r/2, i = 1, 2, . . . , N }
is open, contained in B(w, r) and w ∈ A(w, r, N ). Hence the sets A(w, r, N )
form a base in s. Thus it remains to prove that Φ−1 (A(w, r, N )) is open in
the weak-∗ topology. By the definition of the weak-∗ topology the functions
x∗ 7→ hx∗ , xi i are continuous, so the sets
{x∗ ∈ X ∗ : |hx∗ , xi i − wi | < r/2}
are open as preimages of B(wi , r/2) ⊂ C and hence
N
\
Φ−1 (A(w, r, N )) = {x∗ ∈ X ∗ : |hx∗ , xi i − wi | < r/2}
i=1
is open. 2
FUNCTIONAL ANALYSIS 131
Since the closed unit ball in an infinitely dimensional space is not compact
(Corollary 3.7) we have
Proposition 16.2. If dim X = ∞, then the identity mapping id : X → X
is not compact.
Theorem 16.3. If An ∈ B(X, Y ) is a sequence of compact operators be-
tween Banach spaces and An → A in norm, then A is compact.
Proof. We only need to prove the implication from left to right. We can
assume that dim R(A) = ∞ as otherwise the claim is obvious. Since A is
compact, R(A) is a union of countably many relatively compact sets, so
R(A) is separable. Let {ϕi }∞
i=1 be an orthonormal basis in R(A) and let
so
dim N (B) ⊕ Z = dim(X/X1 ) = dim N (A) ,
Proof. First we will prove that dim N (I + K) < ∞. To this end it suffices
to prove that the closed unit ball in N (I + K) is compact, see Theorem 3.7.
Let xn ∈ N (I + K), kxn k ≤ 1. Since K is compact, after passing to a
subsequence we may assume that Kxn → y converges. Since xn + Kxn = 0,
xn → −y.
Now we will prove that R(I + K) is closed. The space N (I + K) is finitely
dimensional, so it is complemented
(16.6) X = N (I + K) ⊕ V
for some closed subspace V ⊂ X. The mapping I + K restricted to V is a
bijection onto R(I + K) and to prove that R(I + K) is closed it suffices to
show that the inverse mapping is continuous
−1
(I + K)|V : R(I + K) → V .
We need to prove continuity at 0 only. Suppose the mapping is not contin-
uous at 0. Then there is a sequence xn ∈ V such that (I + K)xn → 0, but
xn 6→ 0. Without loss of generality we may assume that kxn k ≥ ε. Then the
sequence yn = xn /kxn k satisfies kyn k = 1, yn + Kyn → 0. By compactness
of K, after passing to a subsequence we have Kyn → y, yn → −y. Hence
kyk = 1, y ∈ V , y ∈ N (I + K) which is a contradiction with (16.6).
FUNCTIONAL ANALYSIS 137
This shows that the sequence {Kxn } cannot have a convergent subsequence,
which contradicts compactness of K. Hence R(I +K) has finite codimension.
We proves that I + K ∈ Fred (X). Now [0, 1] 3 t 7→ I + tK is a continuous
family of Fredholm operators and from the previous result we have ind (I +
K) = ind I = 0. 2
Definition. Let A, B ∈ B(X, Y ). We say that A is congruent to B modulo
compact operators if A − B is compact and we write
A≡B mod K(X, Y ) .
It is easy to see that this is an equivalence relation. Moreover if
A≡B mod K(X, Y ), A1 ≡ B1 mod K(Y, Z) ,
then
A1 A ≡ B1 B mod K(X, Z) .
Indeed
A1 A − B1 B = A1 (A − B) + (A1 − B1 )B
is compact by Theorem 16.5.
Definition. We say that A ∈ B(X, Y ) is invertible modulo compact opera-
tors if there is A1 ∈ B(Y, X) such that
AA1 ≡ IY mod K(Y ), A1 A ≡ IX mod K(X) ,
i.e. A1 A = IX + K1 and AA1 = IY + K2 for some K1 ∈ K(X), K2 ∈ K(Y ).
We call A1 inverse of A modulo compact operators.
Theorem 16.11. Let A ∈ B(X, Y ) be a bounded operator between Banach
spaces. Then A is Fredholm if and only if it is invertible modulo compact
operators.
Proof.51 It suffices to prove that for any r > 0 the number of eigenvalues
satisfying |λ| ≥ r is finite. If this is not true, then there are distinct eigenval-
ues λi , |λi | ≥ r, i = 1, 2, 3, . . . and corresponding eigenvectors xi , kxi k = 1.
Let
Hn = span {x1 , . . . , xn } .
By the Riesz lemma (Theorem 3.6) we can find wn ∈ Hn such that
1
kwn k = 1, kwn − yk ≥ for any y ∈ Hn−1 .
2
We have wn = an xn +yn−1 , yn−1 ∈ Hn−1 Thus for k < n, Awk ∈ Hk ⊂ Hn−1
and hence
kAwn − Awk k = kan λn xn + Ayn−1 − Awk k
r
= |λn |k an xn + yn−1 −(yn−1 − λ−1
n (Ayn−1 − Awk ))k ≥ .
| {z } | {z } 2
wn ∈Hn−1
The first part of the following definition already appears in Section 4.2.
Definition. Let X be a normed space. For linear subspaces M ⊂ X, N ⊂
X ∗ we define annihilators
M ⊥ = {x∗ ∈ X ∗ : hx∗ , xi = 0 for all x ∈ M } ,
⊥
N = {x ∈ X : hx∗ , xi = 0 for all x∗ ∈ N } .
Clearly M ⊥ and ⊥N are closed subspaces of X ∗ and X respectively.
The following result provides a complete description of solvability of the
equations (16.8) and (16.9).
Theorem 16.18 (Riesz-Schauder). If X is a Banach space, A ∈ B(X) is
compact and λ 6= 0, then
(16.10) dim N (A − λI) = dim N (A∗ − λI) < ∞ .
Moreover
(16.11) R(A − λI) =⊥N (A∗ − λI), R(A∗ − λI) = N (A − λI)⊥ .
Before we prove the theorem we will see how it applies to the equations
(16.8) and (16.9). The following description is called the Fredholm alterna-
tive.
It follows from (16.10) that 0 6= λ ∈ C is not an eigenvalue of A if an only
if it is not an eigenvalue of A∗ and in this case both equations
Ax − λx = y, A∗ x∗ − λx∗ = y ∗
have unique solutions for all y ∈ X and y ∗ ∈ X ∗ .
Also 0 6= λ ∈ C is an eigenvalue of A if and only if it is an eigenvalue of
A∗ and the dimensions corresponding of eigenspaces for A and A∗ are finite
and equal. Moreover the equation
Ax − λx = y
has a solution if and only if y ∈⊥N (A∗ − λI), i.e. if
hy ∗ , yi = 0 whenever A∗ y ∗ = λy ∗ .
Note that since dim(A∗ − λI) < ∞, this is a finite number of conditions to
check. Similarly the equation
A∗ x∗ − λx∗ = y ∗
has a solution if and only if y ∗ ∈ N (A − λI)⊥ , i.e.
hy ∗ , yi = 0 whenever Ay = λy.
142 PIOTR HAJLASZ
The above theory was first considered by Fredholm in the setting of inte-
gral equations. Let Ω ⊂ Rn be an open set and let K ∈ L2 (Ω × Ω). As we
know the operator K : L2 (Ω) → L2 (Ω),
Z
Kf (x) = K(x, y)f (y) dy
Ω
(a) The equation (16.12) has a solution for every g ∈ L2 (Ω) if and only
if the only solution to
Z
(16.13) f (x) − µ K(x, y)f (y) dy = 0
Ω
is f = 0.
(b) The equation (16.13) has nonzero solutions, if and only if there are
nonzero solutions to
Z
(16.14) f (y) − µ K(x, y)f (x) dx = 0
Ω
Recall that
⊥
N (T ∗ ) = {y ∈ Y : hy ∗ , yi = 0 whenever T ∗ y ∗ = 0} .
Note that this functional is well defined, because if Ax = Ax̃, then hx1 , xi =
hx∗1 , x̃i by (16.16). The functional Λ can be extended linearly to a functional54
on Cn−1 , so it is of the form
n
X
hΛ, [y2 , . . . , yn ]i = aj yj
j=2
54Linear algebra.
FUNCTIONAL ANALYSIS 147
∗
since yn+1 ∈ N (A∗ − λI) and hyn+1
∗ , y i = 0 for i = 1, 2, . . . , n. The contra-
i
diction proves that n ≥ m, i.e.
dim N (A − λI) ≥ dim N (A∗ − λI) .
Applying this result to A∗ in place of A we have
dim N (A∗ − λI) ≥ dim N (A∗∗ − λI) .
If x ∈ N (A − λI), then it easily follows that κ(x) ∈ N (A∗∗ − λI) and hence
dim N (A∗∗ − λI) ≥ dim N (A − λI) ,
so the above inequalities give (16.10).
Proof. Since hAx, xi = hx, Axi = hAx, xi, (a) follows. If T x = λx, x 6= 0,
then
λhx, xi = hT x, xi = hx, T xi = λhx, xi
so λ ∈ R which is (b). If λ1 6= λ2 are eigenvalues and x1 , x2 6= 0 are corre-
sponding eigenvectors, then using the fact that λ2 ∈ R we have
λ1 hx1 , x2 i = hT x1 , x2 i = hx1 , T x2 i = λ2 hx1 , x2 i ,
so hx1 , x2 i = 0 which is (c). 2
The spectrum of a bounded operator is bounded. more precisely if λ ∈
σ(T ), then |λ| ≤ kT k (see a remark proceeding Theorem 16.14).
Theorem 16.25. If A ∈ B(H) is compact and self-adjoint, then kAk or
−kAk is an eigenvalue of A.
Hence
kA2 yk kyk ≥ hA2 y, yi = kAyk2 ≥ kAk4 = kAk2 kyk2 ≥ kA2 yk kyk ,
so
hA2 y, yi = kA2 yk kyk.
This is possible only if the vectors A2 y and y are parallel, A2 y = cy. We
have
hA2 y, yi kAk4
c= = = kAk2 .
hy, yi kAk2
Let x = y + kAk−1 Ay. If x = 0, then Ay = −kAk y and hence −kAk is an
eigenvalue of A. If x 6= 0, then
Ax = Ay + kAk−1 A2 y = Ay + kAk−1 kAk2 y = Ay + kAk y = kAk x ,
so kAk is an eigenvalue. 2
Let A ∈ B(H) be compact and self-adjoint. According to Theorem 16.16
all nonzero eigenvalues form a finite or infinite sequence {λi }N
i=1 (N is finite
or N = ∞). If N = ∞ and we order the eigenvalues in such a way that
|λ1 | ≥ |λ2 | ≥ |λ3 | ≥ . . ., then
lim λi = 0.
i→∞
Let
Eλi = {x ∈ H : Ax = λi x} .
Eλi is the eigenspace corresponding to the eigenvalue λi . Since Eλi = N (A−
λi I),
dim Eλi < ∞.
Moreover Theorem 16.24 yields
Eλi ⊥ Eλj for i 6= j.
Theorem 16.26 (Spectral theorem). Let A ∈ B(H) be compact and self-
adjoint. Then
N
M
H = ker A ⊕ Eλi .
i=1
Proof. Let
N
M
Y = Eλi .
i=1
Then H = Y ⊕ Y ⊥ and it remains to show that Y ⊥ = ker A. It is easy to
see that A(Y ) ⊂ Y and hence for every x ∈ Y ⊥ and y ∈ Y we have
hAx, yi = hx, Ayi = 0 ,
i.e. Ax ∈ Y ⊥ . Hence A|Y ⊥ : Y ⊥ → Y ⊥ is a compact self-adjoint operator.
According to Theorem 16.25 at least one of the numbers ±kA|Y ⊥ k is its
eigenvalue. It cannot be a nonzero eigenvalue, because all eigenvectors with
FUNCTIONAL ANALYSIS 149
55Laplace-Beltrami.
150 PIOTR HAJLASZ
Proof. If {uk } ⊂ W m,p (Ω) is a Cauchy sequence, then for every α, |α| ≤ m,
{Dα uk } is a Cauchy sequence in Lp (Ω), so Dα uk converges to some function
uα ∈ Lp (Ω) (we will writeu instead of u0 ). Since
Z Z Z Z
α α |α| α |α|
uD ϕ ← uk D ϕ = (−1) D uk ϕ → (−1) uα ϕ
Ω Ω Ω Ω
we conclude that α
D u = uα weakly, so u ∈ W m,p (Ω) and uk → u in the
norm of W m,p (Ω). 2
Theorem 16.28 (Meyers-Serrin). For 1 ≤ p < ∞ smooth functions are
dense in W m,p (Ω), i.e. for every u ∈ W m,p (Ω) there exist a sequence uk ∈
C ∞ (Ω), kuk km,p < ∞ such that ku − uk km,p → 0 as k → ∞.
We will not prove it. The space of functions u ∈ C ∞ (Ω) with kukm,p <
inf ty forms a normed space with respect to the Sobolev norm and the
above theorem shows that the Sobolev space can be equivalently defined as
its completion.
Theorem 16.29. For 1 < p < ∞ the Sobolev space W m,p (Ω) us reflexive.
As before P u is unique (if it exists) and coincides with the classical dif-
ferential operator applied to u is u is smooth.
Definition. For 1 ≤ p < ∞ and m ≥ 1 we define W0m,p (Ω) as the closure
of C0∞ (Ω) in the Sobolev norm.
Theorem 16.30 (Poincaré lemma). If Ω ⊂ Rn is bounded, 1 ≤ p < ∞ and
m ≥ 1, then there is a constant C = C(m, p, Ω) such that
kDα ukp for all u ∈ W0m,p (Ω).
X
kukm,p ≤ C
|α|=m
Proof. We are left with the proof that T is self-adjoint. Let g1 , g2 ∈ L2 (Ω)
and f1 = T g1 , f2 = T g2 . We have to prove that
hT g1 , g2 i = hg1 , T g2 i ,
i.e.
Z Z
(16.19) f1 g2 = g1 f2 .
Ω Ω
Since f1 = T g1 is a weak solution to −∆f1 = g1 we have
Z Z
(16.20) ∇f1 · ∇h = g1 h for h ∈ W01,2 (Ω)
Ω Ω
and similarly −∆f2 = g2 yields
Z Z
(16.21) ∇f2 · ∇h = g2 h for h ∈ W01,2 (Ω).
Ω Ω
The two equalities give
Z Z Z Z
(16.21) (16.20)
f1 g2 = ∇f2 · ∇f1 = ∇f2 ·∇f1 = g1 f2
Ω |{z} Ω |{z} Ω |{z} Ω |{z}
h ∇h ∇h̃ h̃
which is (16.19). 2
Since the operator T : L2 → L2 is compact and self-adjoint we can apply
the spectral theorem.
Definition. Let Ω ⊂ Rn be open and bounded. We say that λ is an eigen-
value of −∆ in W01,2 (Ω) if there is 0 6= g ∈ W01,2 (Ω) such that
−∆g = λg weakly in Ω.
Lemma 16.34. If Ω ⊂ Rn is bounded, the eigenvalues of −∆ in W01,2 (Ω)
are strictly positive.
As we are already observed the eigenspaces Eµi for T are equal to the
eigenspaces Eµ−1 for −∆, so we have
i
(a) The eigenvalues of the Laplace operator −∆ : W01,2 (Ω) → L2 (Ω) are
positive and form an increasing sequence
λ1 < λ2 < λ3 < . . . , lim λi = ∞ .
i→∞
(a) e − x is invertible.
(b)
kxk2
k(e − x)−1 − e − xk ≤ .
1 − kxk
(c) |φ(x)| < 1 for every complex homomorphism φ on A.
Proof. Proof of the part (a) is very similar to that of Theorem 2.7. The
series
X∞
xn = e + x + x2 + x3 + . . .
n=0
converges absolutely, because kxn k ≤ kxkn and kxk < 1. Hence it converges
to an element in A. It is easy to see that this element is an inverse of e − x.
Since
kxk2
k(e − x)−1 − e − xk = kx2 + x3 + . . . k ≤
1 − kxk
the part (b) follows. Finally if λ ∈ C, |λ| ≥ 1, then e − λ−1 x is invertible by
(a) and thus
1 − λ−1 φ(x) = φ(e − λ−1 x) 6= 0
by Proposition 17.3. Hence φ(x) 6= λ for any such λ, so |φ(x)| < 1. 2
Corollary 17.5. Every complex homomorphism on a Banach algebra is
continuous.
158 PIOTR HAJLASZ
Thus we are left with the proof of the property (17.5). The above argu-
ments were purely algebraic and would work in any complex algebra with
unit, however, the proof of (17.5) is analytic.
By the assumptions, there are no invertible elements in N , so ke − xk ≥ 1
for every x ∈ N by Theorem 17.4. Hence
kλe − xk ≥ |λ| = φ(λe − x) for x ∈ N .
Since every element in A is of the form λe−x, x ∈ N (see (17.2) we conclude
that φ is a continuous linear functional of norm 1.
Lemma 17.8. Suppose f is an entire function on C such that f (0) = 1,
f 0 (0) = 0 and
0 < |f (λ)| ≤ e|λ| for λ ∈ C.
Then f (λ) = 1 for all λ ∈ C.
Proof. Since f has no zeroes, there is another entire function g such that
f = eg .
Clearly g(0) = g 0 (0) = 0 and
(17.7) re g(λ) ≤ |λ|.
If |λ| ≤ r, then re g(λ) ≤ r and hence
|re g(λ)| ≤ |2r − re g(λ)| ,
so
|g(λ)|2 = (re g(λ))2 +(im g(λ))2 ≤ (2r −re g(λ))2 +(im g(λ))2 = |2r −g(λ)|2 .
Thus
(17.8) |g(λ)| ≤ |2r − g(λ)| for |λ| ≤ r.
Consider the function
r2 g(λ)
(17.9) hr (λ) = .
λ2 (2r − g(λ))
This fuinction is holomorphic in the disc {λ : |λ| < 2r} by (17.7) and
|hr (λ)| ≤ 1 if |λ| = r by (17.8). Hence the maximum principle gives
(17.10) |hr (λ)| ≤ 1 for |λ| ≤ r.
if we fix λ and let r → ∞ (17.9) and (17.10) gives g(λ) = 0. Hence f (λ) =
e0 = 1. 2
Now we are ready to complete the proof of (17.5). Let a ∈ N . We can
assume that kak = 1. Since φ has norm 1, |φ(an )| ≤ kan k ≤ kakn = 1 and
hence the function
∞
X φ(an ) n
f (λ) = λ , λ∈C
n!
n=0
160 PIOTR HAJLASZ
Proof. The lemma implies that G(A) is open. It also implies that
k(x + h)−1 − x−1 k ≤ 2kx−1 k3 khk2 + kx−1 k2 khk
FUNCTIONAL ANALYSIS 161
which yields continuity of x 7→ x−1 . Since x 7→ x−1 maps G(A) onto G(A)
and it is its inverse, it follows that x 7→ x−1 is a homeomorphism of G(A)
onto itself. 2
The following lemma is often useful.
Lemma 17.11. let {cn }∞
n=1 be a sequence of nonnegative numbers such that
(17.12) cm+n ≤ cm cn
for all positive integers m, n. Then the limit
lim c1/n
n→∞ n
1/n
exists and equals inf n≥1 cn .
Hence
lim sup kxn k1/n ≤ r for all r > ρ(x),
n→∞
i.e.
(17.18) lim sup kxn k1/n ≤ ρ(x) .
n→∞
FUNCTIONAL ANALYSIS 163
Proof. We will prove that a is invertible as the proof for b is the same.
Since ba(ba)−1 b = bab(ab)−1 we have (ba)−1 b = b(ab)−1 and thus
a[b(ab)−1 ] = e, [b(ab)−1 ]a = (ba)−1 ba = e
so a−1 = b(ab)−1 . 2
Lemma 17.14. If λ ∈ σ(x), then λn ∈ σ(xn ).
Proof. We have
λn e − xn = (λe − x)(λn−1 e + . . . + xn−1 ) = (λn−1 e + . . . + xn−1 )(λe − x) .
If λn 6∈ σ(xn ), then according to the previous lemma λe − x is invertible, so
λ 6∈ σ(x). 2
If λ ∈ σ(x), then λn ∈ σ(xn ) and hence |λ|n = |λn | ≤ kxn k, so ρ(x) ≤
kxn k1/nfor any n = 1, 2, . . ., i.e.
ρ(x) ≤ inf kxn k1/n
n≥1
Proof. (a) Let J ⊂ A be a proper ideal. Consider the family of all proper
ideals that contain J. The family is partially ordered by inclusion. According
to the Hausdorff maximality theorem there is a maximal totally ordered
subfamily. The union of this subfamily is also an ideal. It is proper since the
FUNCTIONAL ANALYSIS 165
unit element does not belong to any of the ideals in the family. Hence the
union is a maximal ideal.
(b) Let M be a maximal ideal. Then M is an ideal. It is proper since M
has no invertible elements and invertible elements form an open set. Hence
M = M by maximality of M . 2
Theorem 18.3. Let A be a commutative Banach algebra, and let ∆ be the
set of all complex homomorphisms of A.
Proof. One can easily check that the functions of the form (18.3) form
a Banach algebra A with respect to the pointwise multiplication and the
supremum norm. For each x ∈ Rn ,
(18.4) A 3 f 7→ f (x)
is a complex homomorphism. If we prove that all complex homomorphisms
are of that form, it will follow that the function f satisfies φ(f ) 6= 0, for all
φ ∈ ∆ and hence f is invertible by Theorem 18.3(c) which is what we want
to prove.
For k = 1, 2, . . . , n put gk (x) = e2πixk , where xk is kth coordinate of
x ∈ Rn . Clearly gk , 1/gk ∈ A and both functions have norm 1. If φ ∈ ∆,
then 1 1
|φ(gk )| ≤ 1 and = φ ≤ 1.
φ(gk ) gk
Hence there is y ∈ Rn such that
(18.5) φ(gk ) = e2πiyk = gk (y), k = 1, 2, . . . , n.
Every trigonometric polynomial P is a linear combination of products of the
functions gk and 1/gk and since φ is a complex homomorphism (18.5) implies
that φ(P ) = P (y). Since φ is continuous and trigonometric polynomials are
58Prove it.
59m · x = m x + . . . + m x .
1 1 n n
FUNCTIONAL ANALYSIS 167
dense in A, we get φ(f ) = f (y) for all f ∈ A and hence φ is of the form
(18.4). 2
Definition. Let ∆ be the set of all complex homomorphisms of a commu-
tative Banach algebra A. The formula
x̂(φ) = φ(x) for φ ∈ ∆
assigns to each x ∈ A a function x̂ : ∆ → C called the Gelfand transform of
x.
Let  be the space of all functions x̂ on ∆. The Gelfand topology on ∆ is
the weakest topology for which all the functions x̂ are continuous. Obviously
 ⊂ C(∆).
Since there is a one-to-one correspondence between ∆ and the maximal
ideals in A, ∆ equipped with the Gelfand topology is called the maximal
ideal space of A.
Theorem 18.6. If A is a commutative Banach algebra, then
Proof. (a) Let A∗ be the dual Banach space and B the closed unit ball in
A∗ .By the Banach-Alaoglu theorem, B is compact in the weak-∗ topology. It
is Hausdorff, because the weak-∗ topology is Hausdorff. Clearly ∆ ⊂ B and
the Gelfand topology is the restriction of the weak-∗ topology to ∆. Thus it
remains to show that ∆ is a closed subset of B in the weak-∗ topology.60
Let φ0 belongs to the closure of ∆ in the weak-∗ topology. We have to
prove that
φ0 (xy) = φ0 (x)φ0 (y), φ0 (e) = 1 .
Fix x, y ∈ A and ε > 0. The set
V = {x∗ ∈ A∗ : |hx∗ , zi i − φ0 (zi )| < ε, i = 1, 2, 3, 4}
where z1 = e, z2 = x, z3 = y, z4 = xy is open in the weak-∗ topology and
φ0 ∈ V. It follows from the definition of the closure that there is φ ∈ ∆ ∩ V,
so |φ(zi ) − φ0 (zi )| < ε for i = 1, 2, 3, 4. In particular
|1 − φ0 (e)| = |φ(e) − φ0 (e)| < ε
gives φ0 (e) = 1. Moreover
φ0 (xy) − φ0 (x)φ0 (y) = (φ0 (xy) − φ(xy)) + (φ(x)φ(y) − φ0 (x)φ0 (y))
= (φ0 (xy) − φ(xy)) + φ(x)(φ(y) − φ0 (y)) + φ0 (y)(φ(x) − φ0 (x))
60The proof of this part is very similar to a corresponding argument in the proof of the
Banach-Alaoglu theorem.
168 PIOTR HAJLASZ
gives
|φ0 (xy) − φ0 (x)φ0 (y)| ≤ (1 + |φ(x)| + |φ0 (y)|) ε
and hence
φ0 (xy) = φ0 (x)φ0 (y) .
(b) λ is in the range of x̂ if x̂(φ) = φ(x) = λ for some φ ∈ ∆ which is
equivalent to λ ∈ σ(x) by Theorem 18.3(e). 2
18.1. C ∗ -algebras.
Definition. Let A be a complex algebra (not necessarily commutative). By
an involution on A we mean a map x 7→ x∗ of A onto itself such that
(x + y)∗ = x∗ + y ∗ , (αx)∗ = αx∗ ,
(xy)∗ = y ∗ x∗ , x∗∗ = x.
If e is a unit element, then one easily verifies that e∗ = e and if x is invertible,
then (x−1 )∗ = (x∗ )−1 .
A Banach algebra A is called a C ∗ -algebra if it is an algebra with involution
that satisfies
kx∗ xk = kxk2 for all x ∈ A .
An element x ∈ A satisfying x = x∗ is called hermitian or self-adjoint.
Note that kxk2 = kx∗ xk ≤ kxkkx∗ k implies kxk ≤ kx∗ k. Hence also
kx∗ k
≤ kx∗∗ k = kxk. Thus
(18.6) kx∗ k = kxk.
Since
|hA∗ Ax, yi| = |hAx, Ayi| ≤ kAk2 kxkkyk
we conclude that kA∗ Ak ≤ kAk2 . On the other hand
kAxk2 = hAx, Axi = hA∗ Ax, xi ≤ kA∗ Akkxk2
implies kAk2 ≤ kA∗ Ak. Thus kA∗ Ak = kAk2 .
Therefore B(H) with the adjoint operator as involution is a C ∗ -algebra.
As a direct consequence of (18.6) we have
kA∗ k = kAk.
Theorem 18.7. Let A be a C ∗ -algebra. If x ∈ A is self-adjoint, then σ(x) ⊂
R and ρ(x) = kxk.
f →
7 f ◦ x̂ ∈ C(∆) is an isometric isomorphism of algebras. Thus for every
f ∈ C(σ(x)) there is unique element Ψf ∈ B such that
(Ψf )ˆ = f ◦ x̂
and hence Ψ defines an isometric isomorphism of C(σ(x)) onto B that sat-
isfies
(18.8) Ψf = (Ψf )∗ for all f ∈ C(σ(x)).
Moreover, if f (λ) = λ on σ(x), then Ψf = x.
Note that such polynomials P (λ) are dense in C(σ(x)). Theorem 18.11 allows
us to define not only polynomial functions of x, but any continuous functions
61By the definition of the Gelfand topology.
62Because ∆ is compact.
63Prove it.
172 PIOTR HAJLASZ
Let A ⊂ B(H) be the closure of the space of all operators of the form
(18.11). Clearly A is a commutative C ∗ -subalgebra of B(H) generated by
T . Note that polynomials (18.10) are dense in C(σ(T )). As a consequence
of Theorem 18.11 we have
Theorem 18.14 (Spectral mapping theorem). The mapping P 7→ P (T )
uniquely extends to the isometric isomorphism of algebras
C(σ(T )) 3 f 7→ f (T ) ∈ A.
Moreover f (T ) = f (T )∗ and σ(f (T )) = f (σ(T )).
Theorem 18.15. If T ∈ B(H) is self-adjoint and σ(T ) ⊂ [0, ∞), then there
is a self-adjoint operator S such that S 2 = T .