Functional Analysis: Gerald Teschl
Functional Analysis: Gerald Teschl
Gerald Teschl
Gerald Teschl
Institut für Mathematik
Nordbergstraße 15
Universität Wien
1090 Wien, Austria
Preface v
Chapter 0. Introduction 1
§0.1. Linear partial differential equations 1
Chapter 1. A first look at Banach and Hilbert spaces 5
§1.1. The Banach space of continuous functions 5
§1.2. The geometry of Hilbert spaces 9
§1.3. Completeness 13
§1.4. Bounded operators 14
Chapter 2. Hilbert spaces 17
§2.1. Orthonormal bases 17
§2.2. The projection theorem and the Riesz lemma 21
§2.3. Orthogonal sums and tensor products 23
§2.4. Compact operators 24
§2.5. The spectral theorem for compact symmetric operators 26
§2.6. Applications to Sturm-Liouville operators 28
Chapter 3. Banach spaces 31
Bibliography 33
Glossary of notations 35
Index 36
Index 37
iii
Preface
The present manuscript was written for my course Functional Analysis given
at the University of Vienna in Winter 2004.
It is available from
http://www.mat.univie.ac.at/~gerald/ftp/book-fa/
Acknowledgments
I’d like to thank ....
Gerald Teschl
Vienna, Austria
October, 2004
v
Chapter 0
Introduction
∂ ∂2
u(t, x) = u(t, x), (0.1)
∂t ∂x2
Here u(t, x) is the temperature distribution at time t at the point x. It
is usually assumed, that the temperature at x = 0 and x = 1 is fixed, say
u(t, 0) = a and u(t, 1) = b. By considering u(t, x) → u(t, x)−a−(b−a)x it is
clearly no restriction to assume a = b = 0. Moreover, the initial temperature
distribution u(0, x) = u0 (x) is assumed to be know as well.
1
2 0. Introduction
Since finding the solution seems at first sight not possible, we could try
to find at least some some solutions of (0.1) first. We could for example make
an ansatz for u(t, x) as a product of two functions, each of which depends
on only one variable, that is,
u(t, x) = w(t)y(x). (0.2)
This ansatz is called separation of variables. Plugging everything into
the heat equation and bringing all t, x dependent terms to the left, right
side, respectively, we obtain
ẇ(t) y ′′ (x)
= . (0.3)
w(t) y(x)
Here the dot refers to differentiation with respect to t and the prime to
differentiation with respect to x.
Now if this equation should hold for all t and x, the quotients must be
equal to a constant −λ. That is, we are lead to the equations
−ẇ(t) = λw(t) (0.4)
and
−y ′′ (x) = λy(x), y(0) = y(1) = 0 (0.5)
which can easily be solved. The first one gives
w(t) = c1 e−λt (0.6)
and the second one
√ √
y(x) = c2 cos( λx) + c3 sin( λx). (0.7)
However, y(x) must also satisfy the boundary conditions y(0) = y(1) = 0.
The first one y(0) = 0 is satisfied if c2 = 0 and the second one yields (c3 can
be absorbed by w(t))
√
sin( λ) = 0, (0.8)
which holds if λ = (πn)2 , n ∈ N. In summary, we obtain the solutions
2
un (t, x) = cn e−(πn) t sin(nπx), n ∈ N. (0.9)
It is not hard to see that with this definition C(I) becomes a normed linear
space:
A normed linear space X is a vector space X over C (or R) with a
real-valued function (the norm) k.k such that
• kf k ≥ 0 for all f ∈ X and kf k = 0 if and only if f = 0,
• kλ f k = |λ| kf k for all λ ∈ C and f ∈ X, and
• kf + gk ≤ kf k + kgk for all f, g ∈ X (triangle inequality).
Once we have a norm, we have a distance d(f, g) = kf − gk and hence
we know when a sequence of vectors fn converges to a vector f . We
will write fn → f or limn→∞ fn = f , as usual, in this case. Moreover, a
mapping F : X → Y between to normed spaces is called continuous if
fn → f implies F (fn ) → F (f ). In fact, it is not hard to see that the norm
is continuous (Problem 1.2).
5
6 1. A first look at Banach and Hilbert spaces
∞
X
kak1 = |aj | (1.2)
j=1
and take m → ∞:
k
X
|aj − anj | ≤ ε. (1.5)
j=1
Since this holds for any finite k we even have ka−an k1 ≤ ε. Hence (a−an ) ∈
ℓ1 (N) and since an ∈ ℓ1 (N) we finally conclude a = an + (−an ) ∈ ℓ1 (N). ⋄
Next we want to know if there is a basis for C(I). In order to have only
countable sums, we would even prefer a countable basis. If such a basis
exists, that is, if there is a set {un } ⊂ X of linearly independent vectors
such that every element f ∈ X can be written as
X
f= cn un , cn ∈ C, (1.10)
n
then the span (the set of all linear combinations) of {un } is dense in X. A
set whose span is dense is called total and if we have a total set, we also
have a countable dense set (consider only linear combinations with rational
coefficients). A normed linear space containing a countable dense set is
called separable. Luckily this is the case for C(I):
Theorem 1.2 (Weierstraß). Let I be a compact interval. Then the set of
polynomials is dense in C(I).
where
1
n!
Z
In = (1 − x2 )n dx = 1 1
−1 2(2 + 1) · · · ( 12 + n)
√ Γ(1 + n)
r
π 1
= π 3 = (1 + O( )). (1.12)
Γ( 2 + n) n n
(Remark: The integral is known as Beta function and the asymptotics follow
from Stirling’s formula.)
Lemma 1.3 (Smoothing). Let un (x) be a sequence of nonnegative continu-
ous functions on [−1, 1] such that
Z Z
un (x)dx = 1 and un (x)dx → 0, δ > 0. (1.13)
|x|≤1 δ≤|x|≤1
(In other words, un has mass one and concentrates near x = 0 as n → ∞.)
Then for every f ∈ C[− 21 , 21 ] which vanishes at the endpoints, f (− 12 ) =
f ( 21 ) = 0, we have that
Z 1/2
fn (x) = un (x − y)f (y)dy (1.14)
−1/2
The same is true for ℓ1 (N), but not for ℓ∞ (N) (Problem 1.4)!
Problem 1.1. Show that |kf k − kgk| ≤ kf − gk.
Problem 1.2. Show that the norm, vector addition, and multiplication by
scalars are continuous. That is, if fn → f , gn → g, and λn → λ then
kfn k → kf k, fn + gn → f + g, and λn gn → λg.
Problem 1.3. Show that ℓ∞ (N) is a Banach space.
Problem 1.4. Show that ℓ1 (N) is separable. Show that ℓ∞ (N) is not sep-
arable (Hint: Consider sequences which take only the value one and zero.
How many are there? What is the distance between two such sequences?).
✒▼❇❇
f❇ f⊥
❇
✶❇
✏
✏✏✏
✏ fk
✏✏
✏
✶
✏✏ u
Proof. It suffices to prove the case kgk = 1. But then the claim follows
from kf k2 = |hg, f i|2 + kf⊥ k2 .
Note that the Cauchy-Schwarz inequality implies that the scalar product
is continuous in both variables, that is, if fn → f and gn → g we have
hfn , gn i → hf, gi.
As another consequence we infer that the map k.k is indeed a norm.
kf + gk2 = kf k2 + hf, gi + hg, f i + kgk2 ≤ (kf k + kgk)2 . (1.27)
But let us return to C(I). Can we find a scalar product which has the
maximum norm as associated norm? Unfortunately the answer is no! The
reason is that the maximum norm does not satisfy the parallelogram law
(Problem 1.7).
Theorem 1.6 (Jordan-von Neumann). A norm is associated with a scalar
product if and only if the parallelogram law
kf + gk2 + kf − gk2 = 2kf k2 + 2kgk2 (1.28)
holds.
In this case the scalar product can be recovered from its norm by virtue
of the polarization identity
1
kf + gk2 − kf − gk2 + ikf − igk2 − ikf + igk2 .
hf, gi = (1.29)
4
Proof. If an inner product space is given, verification of the parallelogram
law and the polarization identity is straight forward (Problem 1.6).
To show the converse, we define
1
kf + gk2 − kf − gk2 + ikf − igk2 − ikf + igk2 .
s(f, g) = (1.30)
4
Then s(f, f ) = kf k2 and s(f, g) = s(g, f )∗ are straightforward to check.
Moreover, another straightforward computation using the parallelogram law
shows
g+h
s(f, g) + s(f, h) = 2s(f, ). (1.31)
2
12 1. A first look at Banach and Hilbert spaces
Note that the parallelogram law and the polarization identity even hold
for skew linear forms.
But how do we define a scalar product on C(I)? One possibility is
Z b
hf, gi = f ∗ (x)g(x)dx. (1.32)
a
The corresponding inner product space is denoted by L2 (I). Note that we
have p
kf k ≤ |b − a|kf k∞ (1.33)
and hence the maximum norm is stronger than the L2 norm.
Suppose we have two norms k.k1 and k.k2 on a space X. Then k.k2 is
said to be stronger than k.k1 if there is a constant m > 0 such that
kf k1 ≤ mkf k2 . (1.34)
It is straightforward to check that
Lemma 1.7. If k.k2 is stronger than k.k1 , then any k.k2 Cauchy sequence
is also a k.k1 Cauchy sequence.
This shows that in infinite dimensional spaces different norms will give
raise to different convergent sequences! In fact, the key to solving prob-
lems in infinite dimensional spaces is often finding the right norm! This is
something which cannot happen in the finite dimensional case.
1.3. Completeness 13
Theorem 1.8. If X is a finite dimensional case, then all norms are equiv-
alent. That is, for given two norms k.k1 and k.k2 there are constants m1
and m2 such that
1
kf k1 ≤ kf k2 ≤ m1 kf k1 . (1.36)
m2
≤ j ≤ n, and assume
Proof. Clearly we can choosePa basis uj , 1P P that k.k2 is
the usual Euclidean norm, k j αj uj k22 = j |αj |2 . Let f = j αj uj , then
by the triangle and Cauchy Schwartz inequalities
X sX
kf k1 ≤ |αj |kuj k1 ≤ kuj k21 kf k2 (1.37)
j j
qP
and we can choose m2 = j kuj k1 .
In particular, if fn is convergent with respect to k.k2 it is also convergent
with respect to k.k1 . Thus k.k1 is continuous with respect to k.k2 and attains
its minimum m > 0 on the unit sphere (which is compact by the Heine-Borel
theorem). Now choose m1 = 1/m.
Problem 1.5. Show that ℓ2 (N) is a separable Hilbert space.
Problem 1.6. Let s(f, g) be a skew linear form and p(f ) = s(f, f ) the
associated quadratic form. Prove the parallelogram law
p(f + g) + p(f − g) = 2p(f ) + 2p(g) (1.38)
and the polarization identity
1
s(f, g) =(p(f + g) − p(f − g) + i p(f − ig) − i p(f + ig)) . (1.39)
4
Problem 1.7. Show that the maximum norm does not satisfy the parallel-
ogram law.
Problem 1.8. Prove the claims made about fn , defined in (1.35), in the
last example.
1.3. Completeness
Since L2 is not complete, how can we obtain a Hilbert space out of it? Well
the answer is simple: take the completion.
If X is a (incomplete) normed space, consider the set of all Cauchy
sequences X̃. Call two Cauchy sequences equivalent if their difference con-
verges to zero and denote by X̄ the set of all equivalence classes. It is easy
to see that X̄ (and X̃) inherit the vector space structure from X. Moreover,
Lemma 1.9. If xn is a Cauchy sequence, then kxn k converges.
14 1. A first look at Banach and Hilbert spaces
is finite.
The set of all bounded linear operators from X to Y is denoted by
L(X, Y ). If X = Y we write L(X, X) = L(X).
Theorem 1.11. The space L(X, Y ) together with the operator norm (1.41)
is a normed space. It is a Banach space if Y is.
Hilbert spaces
17
18 2. Hilbert spaces
with equality holding if and only if f lies in the span of {uj }nj=1 .
Of course, since we cannot assume H to be a finite dimensional vec-
tor space, we need to generalize Lemma 2.1 to arbitrary orthonormal sets
{uj }j∈J . We start by assuming that J is countable. Then Bessel’s inequality
(2.6) shows that
X
|huj , f i|2 (2.7)
j∈J
converges absolutely. Moreover, for any finite subset K ⊂ J we have
X X
k huj , f iuj k2 = |huj , f i|2 (2.8)
j∈K j∈K
P
by the Pythagorean theorem and thus j∈J huj , f iuj is Cauchy if and only
2
P
if j∈J |huj , f i| is. Now let J be arbitrary. Again, Bessel’s inequality
shows that for any given ε > 0 there are at most finitely many j for which
|huj , f i| ≥ ε (namely at most kf k/ε). Hence there are at most countably
many j for which |huj , f i| > 0. Thus it follows that
X
|huj , f i|2 (2.9)
j∈J
Note that from Bessel’s inequality (which of course still holds) it follows
that the map f → fk is continuous.
P particularly interested in the case where every f ∈ H
Of course we are
can be written as j∈J huj , f iuj . In this case we will call the orthonormal
set {uj }j∈J an orthonormal basis.
If H is separable it is easy to construct an orthonormal basis. In fact, if H
is separable, then there exists a countable total set {fj }N j=1 . After throwing
away some vectors we can assume that fn+1 cannot be expressed as a linear
combinations of the vectors f1 , . . . fn . Now we can construct an orthonormal
set as follows: We begin by normalizing f1
f1
u1 = . (2.14)
kf1 k
Next we take f2 and remove the component parallel to u1 and normalize
again
f2 − hu1 , f2 iu0
u2 = . (2.15)
kf2 − hu1 , f2 iu1 k
Proceeding like this we define recursively
fn − n−1
P
j=1 huj , fn iuj
un = Pn−1 . (2.16)
kfn − j=1 huj , fn iuj k
This procedure is known as Gram-Schmidt orthogonalization. Hence
we obtain an orthonormal set {uj }N n n
j=1 such that span{uj }j=1 = span{fj }j=1
for any finite n and thus also for N . Since {fj }N
j=1 is total, we infer that
N
{uj }j=1 is an orthonormal basis.
20 2. Hilbert spaces
Theorem 2.3. Every separable inner product space has a countable or-
thonormal basis.
If fact, if there is one countable basis, then it follows that every other
basis is countable as well.
Theorem 2.4. If H is separable, then every orthonormal basis is countable.
Theorem 2.6. For an orthonormal set {uj }j∈J in an Hilbert space H the
following conditions are equivalent:
(i) {uj }j∈J is a maximal orthogonal set.
(ii) For every vector f ∈ H we have
X
f= huj , f iuj . (2.20)
j∈J
By continuity of the norm it suffices to check (iii), and hence also (ii),
for f in a dense set.
Suppose H and H̃ are two Hilbert spaces. Our goal is to construct their
tensor product. The elements should be products f ⊗ f˜ of elements f ∈ H
and f˜ ∈ H̃. Hence we start with the set of all finite linear combinations of
elements of H × H̃
n
αj (fj , f˜j )|(fj , f˜j ) ∈ H × H̃, αj ∈ C}.
X
F(H, H̃) = { (2.30)
j=1
and we compute
X
hf, f i = |αjk |2 > 0. (2.34)
j,k
The completion of F(H, H̃)/N (H, H̃) with respect to the induced norm is
called the tensor product H ⊗ H̃ of H and H̃.
Lemma 2.10. If uj , ũk are orthonormal bases for H, H̃, respectively, then
uj ⊗ ũk is an orthonormal basis for H ⊗ H̃.
Example. We have H ⊗ Cn = Hn . ⋄
where equality has to be understood in the sense, that both spaces are
unitarily equivalent by virtue of the identification
X∞ ∞
X
( fj ) ⊗ f = fj ⊗ f. (2.36)
j=1 j=1
(0) (1)
Proof. Let fj be a bounded sequence. Chose a subsequence fj such
(1) (1) (1)
that A1 fj converges. From fj choose another subsequence fj such that
(2) (n)
A2 fj converges and so on. Since fj might disappear as n → ∞, we con-
(j)
sider the diagonal sequence fj = fj . By construction, fj is a subsequence
(n)
of fj for j ≥ n and hence An fj is Cauchy for any fixed n. Now
kAfj − fk k = k(A − An )(fj − fk ) + An (fj − fk )k
≤ kA − An kkfj − fk k + kAn fj − An fk k (2.37)
shows that Afj is Cauchy since the first term can be made arbitrary small
by choosing n large and the second by the Cauchy property of An fj .
Proof. First of all note that K(., ..) is continuous on [a, b] × [a, b] and hence
uniformly continuous. In particular, for every ε > 0 we can find a δ > 0
such that |K(y, t) − K(x, t)| ≤ ε whenever |y − x| ≤ δ. Let g(x) = Kf (x),
then
Z b
|g(x) − g(y)| ≤ |K(y, t) − K(x, t)| |f (t)|dt
a
Z b
≤ ε |f (t)|dt ≤ εk1k kf k, (2.39)
a
whenever |y − x| ≤ δ. Hence, if fn (x) is a bounded sequence in L2 (a, b),
then gn (x) = Kfn (x) is equicontinuous and has a uniformly convergent
subsequence by the Arzelà-Ascoli theorem (Theorem 2.15 below). But a
uniformly convergent sequence is also convergent in the norm induced by
the scalar product. Therefore K is compact.
26 2. Hilbert spaces
Note that (almost) the same proof shows that K is compact when defined
on C[a, b].
Theorem 2.15 (Arzelà-Ascoli). Suppose the sequence of functions fn (x),
n ∈ N, on a compact interval is (uniformly) equicontinuous, that is, for
every ε > 0 there is a δ > 0 (independent of n) such that
|fn (x) − fn (y)| ≤ ε if |x − y| < δ. (2.40)
If the sequence fn is bounded, then there is a uniformly convergent subse-
quence.
(The proof is not difficult but I still don’t want to repeat it here since it
is covered in most real analysis courses.)
Compact operators are very similar to (finite) matrices as we will see in
the next section.
Problem 2.1. Show that compact operators form an ideal.
Now we show that A has an eigenvalue at all (which is not clear in the
infinite dimensional case)!
Theorem 2.17. A symmetric compact operator has an eigenvalue α1 which
satisfies |α1 | = kAk.
2.5. The spectral theorem for compact symmetric operators 27
Remark: There are two cases where our procedure might fail to con-
struct an orthonormal basis of eigenvectors. One case is where there is
an infinite number of nonzero eigenvalues. In this case αn never reaches 0
and all eigenvectors corresponding to 0 are missed. In the other case, 0 is
reached, but there might not be a countable basis and hence again some of
the eigenvectors corresponding to 0 are missed. In any case one can show
that by adding vectors from the kernel (which are automatically eigenvec-
tors), one can always extend the eigenvectors uj to an orthonormal basis of
eigenvectors.
This is all we need and it remains to apply these results to Sturm-
Liouville operators.
Problem 2.2. Show that if A is bounded, then every eigenvalue α satisfies
|α| ≤ kAk.
d2
L=− + q(x), (2.53)
dx2
2.6. Applications to Sturm-Liouville operators 29
Proof. Pick a value λ ∈ R such that RL (λ) exists. By Theorem 2.18 there
are eigenvalues αn of RL (λ) with corresponding eigenfunctions un . More-
over, RL (λ)un = αn un is equivalent to Lun = (λ + α1n )un , which shows that
En = λ + α1n are eigenvalues of L with corresponding eigenfunctions un .
Now everything follows from Theorem 2.18 except that the eigenvalues are
simple. To show this, observe that if un and vn are two different eigenfunc-
tions corresponding to En , then un (0) = vn (0) = 0 implies W (un , vn ) = 0
and hence un and vn are linearly dependent.
Problem 2.3. Show that the integral operator
Z 1
(Kf )(x) = K(x, y)f (y)dy, (2.60)
0
where K(x, y) ∈ C([0, 1] × [0, 1]) is symmetric if K(x, y)∗ = K(x, y).
Chapter 3
Banach spaces
31
Bibliography
33
Glossary of notations
35
36 Glossary of notations
I . . . identity operator
√
z . . . square root of z with branch cut along (−∞, 0)
z ∗ . . . complex conjugation
k.k . . . norm
k.kp . . . norm in the Banach space Lp
h., ..i . . . scalar product in H
⊕ . . . orthogonal sum of linear spaces or operators, 23
∂ . . . gradient
∂α . . . partial derivative
M⊥ . . . orthogonal complement, 21
(λ1 , λ2 ) = {λ ∈ R | λ1 < λ < λ2 }, open interval
[λ1 , λ2 ] = {λ ∈ R | λ1 ≤ λ ≤ λ2 }, closed interval
Index
Operator
37