Bedos LinAlg
Bedos LinAlg
Analysis
Erik Bédos
Version 1.1
The first draft of these notes were written during the spring of 2018 and tested
simultaneously as a part of the curriculum for the course MAT3400/4400
at the University of Oslo. Hearthy thanks to Adam Sørensen, who was in
charge of the lectures, for his patience and for his comments.
I have tested myself the notes during the spring of 2019 and used this
opportunity to perform some minor revisions. I am grateful to Ulrik Enstad
and to some of the students following the course for pointing out some
misprints in the text. If you find others, or if you have some suggestions,
I’ll be happy if you send me a message.
E.B.
Contents
Acknowledgements i
Contents ii
1 Preliminaries 1
1.1 Normed spaces . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Inner product spaces . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Linear operators . . . . . . . . . . . . . . . . . . . . . . . . . 5
2 On Lp -spaces 9
2.1 The case 1 ≤ p < ∞ . . . . . . . . . . . . . . . . . . . . . . 9
2.2 The case p = ∞ . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
5 On compact operators 79
Contents
iii
CHAPTER 1
Preliminaries
In this chapter we fix some notation and give a review of some of the concepts
and results that we will need. These are usually covered in undergraduate
courses in real analysis, and the reader may consult the book of T. Lindstrøm,
Spaces: an introduction to real analysis (AMS 2017), or any other standard
book in real analysis, for details and proofs.
Definition 1.1.4. Assume {xn }∞ n=1 is a sequence in a normed space (X, k·k).
We say that the series ∞ n=1 n is convergent in X if there is some x∈X
P
x
such that kx − n=1 xn k → 0 as N → ∞, in which case we say that n=1 xn
PN P∞
d(x, y) = kx − yk
2
1.2. Inner product spaces
3
1. Preliminaries
j∈J
4
1.3. Linear operators
Definition 1.3.2. Assume that X and Y are both normed spaces over F.
Then a linear operator T : X → Y is called bounded if there exists some
real number M > 0 such that
kT (x)k ≤ M kxk ∀ x ∈ X,
Proposition 1.3.3. Assume that X and Y are both normed spaces over F
and let T ∈ L(X, Y ). Then the following conditions are equivalent:
(a) T is bounded.
(c) T is continuous on X.
(d) T is continuous at x = 0.
5
1. Preliminaries
Proposition 1.3.4. Assume that X and Y are both normed spaces over F.
For T ∈ B(X, Y ), set
n o
kT k := sup kT (x)k : x ∈ X1 < ∞.
Then the map T → kT k is a norm on B(X, Y ), called the operator norm.
Moreover, we have
n o
kT k = sup kT (x)k : x ∈ X, kxk = 1 (when X 6= {0}),
and
kT (x)k ≤ kT k kxk ∀ x ∈ X.
Theorem 1.3.5. Assume that X is a normed space over F, while Y is a
Banach space. Then B(X, Y ) is Banach space. In particular, B(X) is a
Banach space whenever X is a Banach space.
An immediate consequence of this theorem is that B(X, F) is a Banach
space whenever X is normed space over F. Elements of L(X, F) are called
linear functionals. Thus B(X, F) consists of the bounded linear functionals
on X; it is usually called the dual space of X and denoted by X ∗ in many
books, or by X ] in others.
Definition 1.3.6. A map T : X → Y between two vector spaces over F is
called a (vector space) isomorphism if T ∈ L(X, Y ) and T is bijective (that
is, T is both one-to-one and onto). It is then easy to check that the inverse
map of T , T −1 : Y → X, is linear, i.e., T −1 ∈ L(Y, X).
Definition 1.3.7. Assume that X and Y are normed spaces over F. A map
T : X → Y is called an isomorphism of normed spaces if T is a (vector
space) isomorphism such that both T and T −1 are bounded.
Definition 1.3.8. Assume that X is a normed space and T ∈ B(X). Then
we say that T is invertible in B(X) if T is an isomorphism of normed spaces.
In other words, an operator T ∈ B(X) is invertible in B(X) if T is bijective
and T −1 ∈ B(X).
Proposition 1.3.9. Let X, Y, Z be normed spaces over F, and let T ∈
B(X, Y ), S ∈ B(Y, Z). Set ST := S ◦ T : X → Z. Then ST ∈ B(X, Z)
and
kST k ≤ kSk kT k .
Corollary 1.3.10. Assume that X is a normed space and S ∈ B(X). For
each n ∈ N, let S n := S · · · S denote the product of S with itself n times,
then S n ∈ B(X) and kS n k ≤ kSkn . Note that by setting S 0 = IX , this
formula also holds when n = 0.
6
1.3. Linear operators
n=0
7
CHAPTER 2
On Lp-spaces
Proof. We first note that if a, b are nonnegative real numbers, then we have
ap bq
ab ≤ + . (2.1.2)
p q
10
2.1. The case 1 ≤ p < ∞
as desired.
11
2. On Lp -spaces
f ∼g ⇔ f =g µ-a.e.
12
2.1. The case 1 ≤ p < ∞
the series ∞n=1 [fn ] is convergent in Lp . It suffices to show that there exists
P
that
N N
hX i N
lim k [fn ] − [F ]kp = lim k fn − F kp = lim k fn − F kp = 0 ,
X X
N →∞ N →∞ N →∞
n=1 n=1 n=1
by
∞
g(x) := |fn (x)| for all x ∈ X.
X
n=1
Clearly, the sequence {gN
p
} of A-measurable nonnegative functions is nonde-
creasing, and it converges pointwise to the A-measurable function g p on X.
Further, using Minkowski’s inequality, we get
N N
k |fn | kp =
X X
kgN kp ≤ kfn kp ≤ S
n=1 n=1
Since g ≥ 0, it follows from [L; Exercise 7.5.6] that g p is finite µ-a.e., hence
p
that g is finite µ-a.e. This means that the series ∞ n=1 fn (x) is absolutely
P
13
2. On Lp -spaces
|f | ≤ M µ-a.e.,
n o
in which case we set kf k∞ := inf M > 0 : |f | ≤ M µ-a.e. .
kgk∞ ≤ kgku .
Indeed, we have µ {x ∈ X : |g(x)| > kgku } = µ(∅) = 0. This gives that
|g| ≤ kgku µ-a.e., and both assertions follow readily.
We note that it may happen that kgk∞ < kgku . For example, consider
the Borel function g on X = R given by g = χ{0} ; letting µ be the Lebsgue
measure on BR , we get
n=1
14
2.2. The case p = ∞
|f | ≤ kf k∞ µ-a.e. (2.2.1)
1
Bn := x ∈ X : |f (x)| > kf k∞ + ∈ A.
n
Hence there must exist at least one N ∈ N such that µ(BN ) > 0. Now, by
definition of kf k∞ , we can find M > 0 such that kf k∞ ≤ M < kf k∞ + N1 and
|f | ≤ M µ-a.e. But this implies that |f | ≤ kf k∞ + N1 µ-a.e., i.e., µ(BN ) = 0,
and we have reached a contradiction.
Proof. Using Lemma 2.2.3 we get that |f g|q = |f |q |g|q ≤ kf kq∞ |g|q µ-a.e.
It follows that Z Z
q q
|f g| dµ ≤ kf k∞ |g|q dµ < ∞ .
X X
inequality.
15
2. On Lp -spaces
that is,
m, n ≥ N ⇒ k fm − fn k∞ < ε . (2.2.2)
For each m, n ∈ N, set
n o
Fm,n := x ∈ X : |fm (x) − fn (x)| > kfm − fn k∞ .
16
2.2. The case p = ∞
m,n∈N m,n∈N
n o
= x ∈ X : |fm (x) − fn (x)| ≤ kfm − fn k∞ for all m, n ∈ N .
Let now ε > 0 be given, and choose N ∈ N such that (2.2.2) holds.
Then for all x ∈ E and all m, n ≥ N , we have
|fm (x) − fn (x)| ≤ k fn − fm k∞ < ε . (2.2.3)
It follows that {fn (x)}n∈N is a Cauchy sequence in C for each x ∈ E. Since
C is complete, this implies that {fn (x)}n∈N is convergent for each x ∈ E,
hence that limn→∞ fn (x) = g(x) for some g(x) ∈ C for each x ∈ E. Thereby
we obtain a function g : E → C, which is AE -measurable since g is the
pointwise limit of the restriction of the fn ’s to E. (Here, AE denotes the
σ-algebra of all sets in A which are contained in E).
We can now extend g to an A-measurable function f : X → C by setting
f (x) = g(x) if x ∈ E, and f (x) = 0 otherwise.
Again, let ε > 0 be given and choose N as above. Then, for all x ∈ E
and all m ∈ N such that m ≥ N , we get from (2.2.3) that
|fm (x) − f (x)| = |fm (x) − g(x)| = lim |fm (x) − fn (x)| ≤ ε .
n→∞
17
2. On Lp -spaces
2.3 Exercises
Exercise 2.1. Let V be a vector space (over F) and let k · k denote a
seminorm on V . Define a relation ∼ on V by setting
v ∼ w ⇔ kv − wk = 0
for v, w ∈ V .
a) Check that ∼ is an equivalence relation.
Denote the equivalence class of v ∈ V by [v], that is,
n o
[v] := w ∈ V : v ∼ w ,
n o
and set Ve := [v] : v ∈ V . Moreover, for v, w ∈ V , and λ ∈ F, set
18
2.3. Exercises
Exercise 2.4. Assume that X = (0, 1], A = the Borel subsets of X and µ
is the Lebesgue measure on A. Let f ∈ M be given by
1
f (x) = √ for all x ∈ (0, 1],
x
and let 1 ≤ p < ∞.
a) Show that f ∈ Lp (X, A, µ) if and only if p < 2, and compute kf kp in
this case.
b) Let ν be the measure on A given by
Z
ν(A) = x dµ(x) for all A ∈ A .
A
19
2. On Lp -spaces
Exercise 2.13. Let p ∈ [1, ∞) and assume that (X, A, µ) is finite, that is,
µ(X) < ∞.
a) Show that L∞ ⊆ Lp .
b) Consider 1 ≤ p ≤ r < ∞ and let f ∈ Lr . Show that f ∈ Lp and
1 1
kf kp ≤ µ(X) p − r kf kr .
Exercise 2.14. Let E denote the space of simple functions in M and let f ∈
L∞ . Show that there exists a sequence {hn } in E such that kf − hn k∞ → 0
as n → ∞. Deduce that the space [E] := {[h] : h ∈ E} is dense in L∞ with
respect to k · k∞ .
20
CHAPTER 3
and with the metric induced by this norm. As we recalled in Section 1.1,
all norms on a finite-dimensional vector space are equivalent. The usual
way to prove this is to consider first Fn and show that any other norm on
Fn is equivalent to k · k2 . A crucial fact in the proof is that a subset of
Fn is compact (w.r.t. the metric associated with k · k2 ) if and only if it is
closed and bounded. It will be useful for us to know that this property,
sometimes called the Heine-Borel property, holds in any finite-dimensional
normed space. We will need the following lemma.
Lemma 3.1.1. Let X and Y be finite-dimensional normed spaces. Assume
that X and Y are isomorphic as vector spaces and let T ∈ L(X, Y ) be an
isomorphism. Then T is an isomorphism of normed spaces.
Proof. We have to show that T and T −1 are bounded. To avoid confusion,
we let k · k and k · k0 denote the respective norms on X and Y . For x ∈ X
set
kxkT := kT (x)k0 .
Clearly, the map x → kxkT is a seminorm on X; in fact, it is a norm since
22
3.1. Aspects of finite dimensionality
Note that the above result is not true in general if we instead assume that
Y is finite-dimensional, even in the case where Y = F : a linear functional
T : X → F may be unbounded when X is an infinite-dimensional normed
space. For an example, see Exercise 3.3.
23
3. More on normed spaces and linear operators
Proposition 3.1.9. Let X and Y be normed spaces over F, and assume that
T ∈ B(X, Y ) has finite-rank. Then, for any given bounded sequence {xn }n∈N
in X, we have that the sequence {T (xn )}n∈N has a convergent subsequence
in Y .
Proof. Assume {xn }n∈N ⊆ X satisfies kxn k ≤ M for all n ∈ N for some
M > 0. Then we have
kT (xn )k ≤ kT k kxn k ≤ kT k M
24
3.2. Direct sums and projections
Definition 3.2.1. We will say that X is the (internal) algebraic direct sum
of M1 and M2 , and write X = M1 +̇ M2 , when
X = M1 + M2 and M1 ∩ M2 = {0} .
(i) X = M1 +̇ M2 ;
Proof. Assume (i) holds and let x ∈ X. Then we have x = x1 + x2 for some
x1 ∈ M1 , x2 ∈ M2 . If we also have x = x01 + x02 for some x01 ∈ M1 , x02 ∈ M2 ,
then we get
x1 − x01 = x02 − x2 ∈ M1 ∩ M2 .
Since M1 ∩ M2 = {0}, this implies that x01 = x1 and x02 = x2 . Thus (ii)
holds.
Conversely, assume (ii) holds. It then obvious that X = M1 + M2 .
Consider y ∈ M1 ∩ M2 . Then we have y = y + 0 with y ∈ M1 , 0 ∈ M2 ,
and y = 0 + y with 0 ∈ M1 , y ∈ M2 . By uniqueness, we get y = 0. Thus,
M1 ∩ M2 = {0}, so (i) holds.
Remark 3.2.3. If V1 and V2 are vector spaces over F, then one may form
their direct product V1 × V2 , which is often called the (external) algebraic
direct sum of V1 and V2 . (This concept is presumably well-known; the
definition is recalled in Exercise 3.5). In the case of an (internal) algebraic
direct sum X = M1 +̇ M2 , it can easily be verified that X is isomorphic to
M1 × M2 .
25
3. More on normed spaces and linear operators
Example 3.2.4. a) Let X be the space of all n × n matrices over F, and let
M1 (resp. M2 ) denote the subspace of X consisting of all upper (resp. lower)
triangular matrices in X. Then it is obvious that we have X = M1 + M2 ;
but X is not the algebraic direct sum of M1 and M2 , since M1 ∩ M2 consists
of all the diagonal matrices.
b) Let X be the space of all n × n matrices over R, and let M1 (resp. M2 )
denote the subspace of symmetric (resp. skew-symmetric) matrices in X.
(We recall that A ∈ X is called skew-symmetric when At = −A.) Then we
have X = M1 +̇ M2 . Indeed, if A ∈ X, then A = A1 + A2 , where
1 1
A1 := (A + At ) ∈ M1 and A2 := (A − At ) ∈ M2 .
2 2
Moreover, if A ∈ M1 ∩ M2 , then we have A = At = −A, so A = 0.
P1 (x) := x1 , P2 (x) := x2 ,
P1 + P2 = I, P1 P2 = P2 P1 = 0,
R2 = M1 +̇ M2
26
3.2. Direct sums and projections
X = P (X) +̇ ker(P ).
Since
P x − P (x) = P (x) − P 2 (x) = 0 ,
27
3. More on normed spaces and linear operators
X = P (X) ⊕ ker(P ) .
V1 × V2 = Ve1 ⊕ Ve2 .
28
3.2. Direct sums and projections
can not be complemented in `∞ (N) (with uniform norm), but we don’t have
yet the tools necessary to prove this. Proposition 3.2.11 tells us that if a
closed subspace M of a normed space X is the range of a projection P in
B(X), then M can be complemented. The previous remark implies that the
converse holds when X is a Banach space. It is also known that a finite
dimensional subspace of a normed space can always be complemented. We
will see in the next chapter that any closed subspace of a Hilbert space can
be complemented (by its orthogonal complement).
Direct sums and projections are useful in connection with the study of
linear operators.
Proposition 3.2.15. Assume X is a vector space such that X = M1 +̇ M2
for some subspaces M1 , M2 . To each S1 ∈ L(M1 ) and S2 ∈ L(M2 ), we may
associate an operator S = S1 +̇S2 ∈ L(X) given by
(S1 +̇S2 )(x) := S1 (x1 ) + S2 (x2 )
for x = x1 + x2 ∈ X with x1 ∈ M1 and x2 ∈ M2 .
If P1 (resp. P2 ) denote the projection from X on M1 along M2 (resp. on
M2 along M1 ), then S = S1 +̇S2 commutes with each Pj , that is, we have
SPj = Pj S for j = 1, 2.
Moreover, if X is a normed space, M1 and M2 are closed in X, and P1
is bounded (or, equivalently, P2 is bounded), then S1 +̇S2 is bounded if and
only if S1 and S2 are bounded.
(Note that if X is a Banach space, then P1 and P2 are automatically
bounded, as mentioned in Remark 3.2.13).
29
3. More on normed spaces and linear operators
Proof. The reader should have no difficulty to provide the necessary details,
so we leave this as an exercise.
30
3.2. Direct sums and projections
Proof. If (a) holds, then it follows from Proposition 3.2.15 that (c) and (d)
hold. Since P2 = I − P1 , it is elementary that (c) is equivalent to (d).
Assume that (c) holds. Let x1 ∈ M1 . Then we have
Finally, we mention for completeness that one can also consider direct
sums decompositions of a vector space with more than two summands.
Let X is a vector space over F, and assume that M1 , M2 , . . . , Mn are
subspaces of X. Then X is said to be the (internal) algebraic direct sum of
M1 , M2 , . . . , Mn if X = M1 + M2 + · · · + Mn and the following independence
condition holds: if x1 ∈ M1 , x2 ∈ M2 , . . . , xn ∈ Mn and
x1 + x2 + · · · + xn = 0 ,
31
3. More on normed spaces and linear operators
as desired.
Since Y is complete, we can conclude that there exists some y ∈ Y
such that limn T0 (xn ) = y. Note that y only depends on x. Indeed, as-
sume {x0n }n∈N is another sequence in X0 converging to x. Then the se-
quence x1 , x01 , x2 , x02 , . . . , xn , x0n , . . . in X0 also converges to x, so, arguing
as above, we get that there exists some z ∈ Y such that the sequence
T0 (x1 ), T0 (x01 ), T0 (x2 ), T0 (x02 ), . . . , T0 (xn ), T0 (x0n ), . . . converges to z. This
implies that
lim
n
T0 (x0n ) = z = lim n
T0 (xn ) = y .
Hence it makes sense to define T (x) := y. Doing this for every x ∈ X, we
get a map T : X → Y , and it is easy to check that T is linear, so we leave
this as an exercise.
32
3.3. Extension by density and continuity
kT (x)k = lim
n
kT0 (xn )k ≤ kT0 k lim
n
kxn k = kT0 k kxk .
33
3. More on normed spaces and linear operators
Example 3.3.6. Let a, b ∈ R, a < b, and equip the space C([a, b]) of all con-
tinuous complex functions on [a, b] with the norm kf k2 = ( ab |f (s)|2 ds)1/2 .
R
Considering the square [a, b] × [a, b] as a metric space w.r.t. the Euclidean
metric inherited from R2 , let K : [a, b] × [a, b] → C be a continuous function.
One can then associate to K an integral operator TK on C([a, b]) as follows.
Let f ∈ C([a, b]). Since the function t → K(s, t) f (t) is continuous on
[a, b] for each s ∈ [a, b], we may define a function TK (f ) : [a, b] → C by
Z b
[TK (f )](s) = K(s, t) f (t) dt for all s ∈ [a, b] .
a
34
3.4. Exercises
3.4 Exercises
Exercise 3.1. Let H be a Hilbert space which is infinite-dimensional (as a
vector space). Argue first that there exists an orthonormal sequence {xn }n∈N
in H. Then use this sequence to show that the unit ball H1 is not compact.
Exercise 3.2. Let X be a normed space. Let M denote a finite-dimensional
subspace of X, and assume M 6= X.
a) Let x ∈ X \ M . Show that d := inf m∈M kx − mk > 0.
b) Show that there exists y ∈ X such that kyk = 1 and
1
≤ ky − mk for all m ∈ M.
2
c) Assume that X is infinite-dimensional (as a vector space). Show that
the unit ball X1 is not compact.
(Hint : Use b) to construct inductively a sequence {yn }n∈N in X1 such
that 1/2 ≤ kyn − yk k for all 1 ≤ k < n.)
Exercise 3.3. Let X be the subspace of `∞ (N) given by
X = {f : N → C : f (n) = 0 for all but finitely many n}.
n=1
35
3. More on normed spaces and linear operators
36
3.4. Exercises
Exercise 3.9. Set X = C2 and let {e1 , e2 } denote the standard basis
of X. Let T ∈ L(X) be the linear operator satisfying T (e1 ) = e1 and
T (e2 ) = i e1 + e2 . Clearly, 1 is an eigenvalue of T . Set M1 = E1T , so M1 is a
subspace of X which is invariant under T .
Show that there is no subspace M2 of X which is invariant under T and
satisfies that X = M1 +̇ M2 .
X = ker(I − S) +̇ ker(I + S)
and that S is the symmetry through ker(I − S) along ker(I + S). Finally,
check that S is decomposable with respect to this direct sum decomposition.
c) Assume now that X is a normed space and that S ∈ B(X) satisfies
S = I. Deduce that X = ker(I − S) ⊕ ker(I + S).
2
d) Let a > 0 and consider the space X = C([−a, a]) with the uniform
norm. Define S : X → X by
37
3. More on normed spaces and linear operators
(x + M ) + (x0 + M ) := (x + x0 ) + M , λ (x + M ) := (λx) + M
for all x, x0 ∈ X and λ ∈ F. You should first argue that these operations
are well-defined.
The map Q : X → X/M given by Q(x) = x + M is called the quotient
map. It is evident that Q is linear.
c) Assume now that X = M +̇ N for some subspaces M and N of X.
Show that X/M is isomorphic to N . (Similarly, X/N is isomorphic to M ).
Hint : Consider the map π : N → X/M given by π := Q|N : N → X/M ,
i.e.,
π(y) := y + M for all y ∈ N , (3.4.1)
and show that π is an isomorphism.
38
3.4. Exercises
(iii) P 0 is bounded.
Exercise 3.14. Consider X = R2 . Find three subspaces M1 , M2 , M3 of X
such that
• X = M1 + M2 + M3 ;
• M1 ∩ M2 ∩ M3 = {(0, 0)};
39
3. More on normed spaces and linear operators
subsets of [a, b] and let K be a continuous complex function on [a, b] × [a, b].
For each s ∈ [a, b], let ks : [a, b] → C denote the continuous function defined
by ks (t) := K(s, t) for all t ∈ [a, b].
b) Let f ∈ L2 ([a, b]) and s ∈ [a, b]. Show the function ks f is Lebesgue
integrable on [a, b] and satisfies
Z
ks f dµ | ≤ kks k2 kf k2 .
[a,b]
Deduce that we obtain a linear map TK0 : L2 ([a, b]) → L2 ([a, b]) by setting
Z
TK0 (f ) (s) := ks f dµ for each f ∈ L2 ([a, b]) and all s ∈ [a, b],
[a,b]
40
CHAPTER 4
1
s ≤ kyn k2 < s + . (4.1.1)
2n
Then the sequence {yn }n∈N is Cauchy in H. Indeed, consider m, n ∈ N.
Then, using the parallellogram law and (4.1.1), we get that
1 1
kyn + ym k2 + kyn − ym k2 = 2 kyn k2 + 2 kym k2 < 4 s + + .
n m
Now, since C is convex, we have c := 12 yn + 21 ym ∈ C. Hence,
kyn + ym k2 = 4 kck2 ≥ 4 s ,
so we get
1 1 1 1
kyn − ym k2 < 4 s + + − kyn + ym k2 ≤ + .
n m n m
42
4.1. Geometry in Hilbert spaces
Thus, given ε > 0, we can choose N ∈ N such that N ≥ (2ε2 )−1 , and obtain
that kyn − ym k < ε for all n, m ≥ N , as desired.
As H is complete, there exists y0 ∈ H such that limn yn = y0 . Since C
is closed, y0 ∈ C. Letting n → ∞ in (4.1.1), we get that
√
ky0 k = s = inf{kyk : y ∈ C}.
If y00 ∈ C also satisfies that ky00 k = inf{kyk : y ∈ C}, then we can consider
the sequence {zn }n∈N in C given by zn = y00 if n is odd and zn = y0 if n is
even. Since zn satisfies (4.1.1) (with yn = zn ) for each n, we can conclude as
above that {zn }n∈N is convergent. This clearly implies that y00 = y0 . Thus,
y0 is the unique vector in C satisfying ky0 k = inf{kyk : y ∈ C}, and we can
set 0C := y0 .
In the general case where x ∈ H, we note that the set
D := {x − y : y ∈ C}
is closed and convex. Using the first part, we get that there exists a unique
vector 0D ∈ D such that k0D k = inf{kzk : z ∈ D} = d(x, C). Then
xC := x − 0D ∈ C has the desired properties.
One important application is when C is a closed subspace M of H.
Theorem 4.1.2. Let M be a closed subspace of a Hilbert space H. Then
we have
H = M ⊕ M⊥ .
The associated projection PM of H on M along M ⊥ is given by
(M ⊥ )⊥ = M and PM ⊥ = IH − PM .
43
4. More on Hilbert spaces
Thus, hx⊥ , yi = 0 in this case too. As this holds for every y ∈ M , the claim
is proven.
Since x = xM + x⊥ , by definition of x⊥ , we get that
H = M + M⊥ .
H = M ⊥ ⊕ (M ⊥ )⊥ .
44
4.1. Geometry in Hilbert spaces
can set D E Z
[f ], [g] := f g dµ .
X
45
4. More on Hilbert spaces
To prove this, assume first that [g] ∈ ME and [h] ∈ MF . Then one easily
sees that g = g 1E µ-a.e. and h = h 1F µ-a.e., so, as E ∩ F = ∅, we get
D E Z Z
[g], [h] = g 1E h 1F dµ = g h 1E∩F dµ = 0 .
X X
Since this is true for all [g] ∈ ME , this implies that [h] ∈ (ME )⊥ . As this
holds for all [h] ∈ MF , we get that MF ⊆ (ME )⊥ .
To show the reverse inclusion, let [h] ∈ (ME )⊥ . Then we have
Z
g h dµ = 0 whenever [g] ∈ ME .
X
46
4.2. Orthonormal bases in Hilbert spaces
[f ] = [f 1E ] + [f 1F ] , where [f 1E ] ∈ ME , [f 1F ] ∈ MF ,
which holds for all [f ] ∈ L2 . From this equation, we now see that the
orthogonal projection of L2 on ME (resp. MF ) is given by
Example 4.2.2. Let H = L2 ([−π, π], A, µ), where A denotes the σ-algebra
of all Lebesgue measurable subsets of [−π, π], and µ is the normalized
Lebesgue measure on A, that is,
1
µ(A) := λ(A) for all A ∈ A ,
2π
where λ denotes the Lebesgue measure on R. In particular, we have
µ([−π, π]) = 1. For each n ∈ Z, let en : [−π, π] → C denote the con-
tinuous function given by
B := {[en ] : n ∈ Z}
47
4. More on Hilbert spaces
48
4.2. Orthonormal bases in Hilbert spaces
E := {ex : x ∈ X} .
y ∈X y ∈{x}
we get that f (x) = 0 for all x ∈ X, i.e., f = 0. This shows that E ⊥ = {0},
and Proposition 4.2.3 gives that E is an orthonormal basis for `2 (X).
49
4. More on Hilbert spaces
n o
i) We set B1 := 1
kx1 k
x1 . Clearly, B1 is an orthonormal basis for M1 .
As we will soon see, a similar expansion also holds in any infinite dimensional
Hilbert space.
We will use the following notation. If j → tj is a function from a
nonempty set J into [0, ∞), then we set
n X o
tj := sup tj : F ⊆ J, F is finite and nonempty ∈ [0, ∞] .
X
j ∈J j ∈F
w.r.t. the counting measure on P(J) (= the σ-algebra of all subsets of J).
We first note that Bessel’s inequality holds for any orthonormal set:
Lemma 4.2.6. Assume that B is an orthonormal set in an inner product
space H, and let x ∈ H. Then
2
≤ kxk2 ,
X
hx, ui
u∈B
n o
and the set Bx := u ∈ B : hx, ui =
6 0 is countable.
50
4.2. Orthonormal bases in Hilbert spaces
The next lemma will be useful at several occasions.
Lemma 4.2.7. Assume {uj : j ∈ N} is a countably infinite orthonormal
set of distinct vectors in a Hilbert space H and let {cj }j∈N be any sequence
in F satisfying that
∞
|cj |2 < ∞ .
X
j=1
j=n+1 j=n+1
Using the assumption, the sum above can be made as small as we want by
choosing m and n large enough. Thus the sequence {yn }n∈N is Cauchy in
H, so it converges to some y ∈ H, i.e., we have
∞
y=
X
cj uj .
j=1
For each k ∈ N, continuity and linearity of the inner product in the first
variable gives then that
∞
hy, uk i = cj huj , uk i = ck .
X
j=1
51
4. More on Hilbert spaces
x= (4.2.4)
X
hx, ui u
u ∈ Bx
where Bx = {u ∈ B : hx, ui =
6 0} is countable (cf. Lemma 4.2.6) and
nonempty.
By (4.2.4) we mean that if Bx is not finite, and Bx = {uj : j ∈ N} is
any enumeration of the distinct elements of Bx , then we have
n ∞
lim x − hx, uj i uj = 0 , i.e., x=
X X
n→∞
hx, uj i uj .
j=1 j=1
2
(c) For every x ∈ H we have kxk2 = hx, ui .
P
u∈B
j=1
Applying Lemma 4.2.7 with cj = hx, uj i for every j ∈ N, we get that the
series ∞j=1 hx, uj i uj converges to some y ∈ H, which satisfies that
P
j=1
It follows that x − y ∈ B = {0}, hence that x = y. This shows that the
⊥
52
4.2. Orthonormal bases in Hilbert spaces
(b) ⇒ (c): Assume (b) holds, and let x ∈ H \ {0}. Again we consider
the more difficult case where Bx is countably infinite, so Bx = {uj : j ∈ N}
as above. By continuity of the norm and Pythagoras’ identity, we get
∞
2
kxk2 =
X
hx, uj i .
j=1
2
Hence, given ε > 0, we can find n ∈ N such that kxk2 −
Pn
j=1 hx, uj i < ε,
giving
n
2 2
kxk2 − ε <
X X
hx, uj i ≤ hx, ui .
j=1 u∈B
2
Since this holds for every ε > 0, we get that kxk2 ≤ u ∈ B hx, ui . Com-
P
bining this inequality with Lemma 4.2.6, we see that (c) holds.
2
(c) ⇒ (a): Assume kxk2 = u ∈ B hx, ui for every x ∈ H. If x ∈ B ⊥ ,
P
53
4. More on Hilbert spaces
PM (x) =
X
hx, vi v ,
v ∈ Cx
for all x ∈ H.
Example 4.2.12. Let B = {[en ] : n ∈ Z} denote the orthonormal basis
for H = L2 ([−π, π], A, λ/2π) described in Example 4.2.2. For [f ] ∈ H and
n ∈ Z it is common to set
1 Z
[f ] (n)
d := h [f ], [en ] i = f (t)e−int dλ(t) ,
2π [−π,π]
which is called the Fourier coefficient of [f ] at n.
In fact, it is usual to write f instead of [f ], having in mind that one
identifies functions which agree µ-a.e. Hence, the Fourier coefficient of f
at n is denoted by fb(n), and the Fourier expansion of f w.r.t. B is then
written as
f= fb(n) en ,
X
n∈Z
meaning that
m
f = lim fb(n) en (w.r.t. k · k2 ).
X
m→∞
n=−m
n∈Z
54
4.3. Adjoint operators
Theorem 4.3.1. Let H be a Hilbert space (over F). For each y ∈ H, define
ϕy : H → F by
ϕy (x) := hx, yi for all x ∈ H .
Then ϕy ∈ H ∗ for all y ∈ H.
Moreover, the map y → ϕy is a bijection from H onto H ∗ , which is
isometric, and conjugate-linear in the sense that
Proof. Let y ∈ H. Then the map ϕy is clearly linear. Moreover, for all
x ∈ H, we have
|ϕy (x)| = |hx, yi| ≤ kxk kyk .
Hence, ϕy is bounded, with kϕy k ≤ kyk. If y 6= 0, then
1 1
ϕy y = hy, yi = kyk,
kyk kyk
y := ϕ(z) z ∈ H .
We claim that ϕ = ϕy . Indeed, let x ∈ H and set m := ϕ(x) z − ϕ(z) x ∈
H. Then we have
55
4. More on Hilbert spaces
for all x, y ∈ H .
The ∗-operation on B(H), T → T ∗ , enjoys the following properties:
For all S, T ∈ B(H) and all α, β ∈ F, we have
• i) (α S + β T )∗ = α S ∗ + β T ∗ ; ii) (ST )∗ = T ∗ S ∗ ; iii) (T ∗ )∗ = T ;
• iv) kT ∗ k = kT k ; v) kT ∗ T k = kT k2 .
Remark 4.3.3. If H and K are Hilbert spaces (over the same F), then one
may associate to each T ∈ B(H, K) a unique adjoint operator T ∗ ∈ B(K, H)
satisfying (4.3.1) for all x ∈ H and all y ∈ K, and enjoying similar properties.
We leave this as an exercise.
Proof of Theorem 4.3.2. Let T ∈ B(H) and consider y ∈ H. Using the
linearity of T and the linearity of the inner product in the first variable, we
get that the map ϕ : H → F defined by
D E
ϕ(x) := T (x), y for all x ∈ H ,
56
4.3. Adjoint operators
57
4. More on Hilbert spaces
58
4.3. Adjoint operators
We may think of the map A sending each (j, k) to A(j, k) as the (infinite)
matrix of T (w.r.t. B) since, for each k ∈ N, we have
∞ D E ∞
T (uk ) = T (uk ), uj uj = (4.3.4)
X X
A(j, k) uj .
j=1 j=1
A∗ (j, k) := A(k, j) ,
j=1
M := sup{|λj | : j ∈ N} < ∞.
59
4. More on Hilbert spaces
j=1
It follows now readily that the map x → D(x) gives an operator D ∈ B(H)
such that kDk ≤ M and satisfying (4.3.5). Since kDk ≥ kD(uk )k = |λk |
for all k ∈ N, we also have that kDk ≥ M . Hence, kDk = M .
It is now obvious that the matrix of D (w.r.t. B) is the diagonal (infinite)
matrix Λ defined for each (j, k) ∈ N by
λ
j if j = k,
Λ(j, k) =
0 otherwise.
The operator D is often called the diagonal operator associated to {λj }j∈N
(w.r.t. B).
From our discussion in the first part, we get that the matrix of D∗ is Λ∗ .
Thus we have D∗ (uk ) = λk uk for all k ∈ N, so D∗ is the diagonal operator
associated to { λj }j∈N (w.r.t. B).
b) We may also easily argue that there exists an operator S ∈ B(H)
satisfying that
S(uk ) = uk+1 for all k ∈ N. (4.3.6)
Indeed, since ∞
n=2 |hx, un−1 i| =
2
j=1 |hx, uj i| = kxk < ∞ for all x ∈ H,
∞ 2 2
P P
60
4.3. Adjoint operators
for each (j, k) ∈ N. Thus, the matrix of S ∗ (w.r.t. B) is the matrix σ ∗ given
by
1 if k = j + 1,
σ ∗ (j, k) = σ(k, j) =
0 otherwise.
for all j, k ∈ N, so we get that
∞ 0 if k = 1,
S ∗ (uk ) = σ ∗ (j, k) uj =
X
j=1 uk−1 if k ≥ 2.
The operator S ∗ is called the left shift operator on H (w.r.t. B). We note
that S ∗ is not isometric, in fact not even injective, since S ∗ (u1 ) = 0.
Indeed, this follows readily from Proposition 2.2.4 (with q = 2). Now, for
all [g], [h] ∈ H, we have
D E Z Z D E
Mf ([g]), [h] = f g h dµ = g f h dµ = [g], Mf ([h]) .
X X
61
4. More on Hilbert spaces
62
4.4. Self-adjoint operators
Example 4.4.4. Let (X, A, µ) be a measure space and set H := L2 (X, A, µ).
If f ∈ L∞ , then the multiplication operator Mf ∈ B(H) defined in Example
4.3.6 is self-adjoint if and only if Mf = Mf .
Thus, Mf is self-adjoint whenever f is real-valued (µ-a.e.). It can be
shown that the converse statement holds whenever (X, A, µ) satisfies the
mild assumption that it is semifinite (cf. Exercise 4.22).
63
4. More on Hilbert spaces
T = Re(T ) + i Im(T ) .
Proof. The first assertion follows readily from Proposition 4.4.7. The second
one is elementary.
Consider a bounded operator T on a Hilbert space H 6= {0}. The
numerical range of T is defined as the subset of F given by
nD E o
WT := T (x), x : x ∈ H, kxk = 1 .
64
4.4. Self-adjoint operators
Combining (4.4.2) and (4.4.3), and using our previous observations, as well
as the parallellogram law and the fact that kxk ≤ 1, kyk = 1, we get
1 D E D E
kT (x)k = T (x), y + T (y), x
2
1 D E D E
= T (x + y), x + y − T (x − y), x − y
4
1
≤ NT kx + yk2 + kx − yk2
4
1
= NT kxk2 + kyk2
2
≤ NT .
65
4. More on Hilbert spaces
Example 4.4.10. Let H = L2 ([0, 1]) (with usual Lebesgue measure) and
let T = Mf be the self-adjoint operator in B(H) given by multiplication with
the bounded continuous function f (t) = t on [0, 1], cf. Example 4.4.4. Then
the reader should have no trouble in checking that T has no eigenvalues.
We will see in the next chapter that every compact self-adjoint operator
can be diagonalized in the sense mentioned above. Theorem 4.4.9 will help us
to make the first step in proving this, by showing that a compact self-adjoint
operator T has at least one an eigenvalue, namely kT k or −kT k.
so (ii) holds.
(ii) ⇒ (iii): Any map preserving the inner product is isometric, so this is
evident.
(iii) ⇒ (i): Assume S is a linear isometry. Then S ∈ B(H) and T :=
S ∗ S − IH ∈ B(H) is self-adjoint. Then for any x ∈ H, we have
D E D E D E
T (x), x = (S ∗ S −I)(x), x = S(x), S(x) −hx, xi = kS(x)k2 −kxk2 = 0
66
4.5. Unitary operators
U ∗ U = U U ∗ = IH .
67
4. More on Hilbert spaces
Example 4.5.7. Let (X, A, µ) be a measure space and set H := L2 (X, A, µ).
For f ∈ L∞ , consider the multiplication operator Mf ∈ B(H). Then we
clearly have
(Mf )∗ Mf = M|f |2 = Mf (Mf )∗ ,
so we see that Mf is unitary whenever |f | = 1 µ-a.e. The converse holds if
µ is semifinite, cf. Exercise 4.29.
Example 4.5.8. Let H = `2 (Z). We may then define the bilateral forward
shift operator U : H → H by
Let now H, K be Hilbert spaces (over F). A bijective, linear map U from
H onto K which preserves the inner product is often called an isomorphism
of Hilbert spaces. As in Proposition 4.5.5, one shows that it is equivalent to
require that U is a surjective linear isometry, or that U ∈ B(H, K) is unitary
in the sense that we have U ∗ U = IH and U U ∗ = IK . (Here, U ∗ ∈ B(K, H)
denotes the adjoint of U , cf. Remark 4.3.3). We will say therefore say that
H and K are isomorphic as Hilbert spaces when such a map U : H → K
exists.
68
4.5. Unitary operators
have xb ∈ `2 (B) and kxbk = kxk. Thus we can define an isometric map
U : H → `2 (B) by
U (x) = xb for all x ∈ H.
It is elementary to check that U is linear. Moreover, U is surjective.
Indeed, let ξ ∈ `2 (B). As u∈B |ξ(u)|2 < ∞, the set
P
Bξ := {u ∈ B : ξ(u) 6= 0}
j∈N
0 if u ∈ B \ Bξ ,
i.e., xb(u) = ξ(u) for all u ∈ B. Hence, U (x) = ξ, showing that U is surjective.
We can now conclude that U is an isomorphism of Hilbert spaces from
H to `2 (B), as we wanted to show.
u∈B
69
4. More on Hilbert spaces
4.6 Exercises
In the exercises of this chapter, H always denotes a Hilbert space over F,
unless otherwise stated.
Exercise 4.6. Consider H = L2 ([a, b], A, µ), where A denotes the σ-algebra
of all Lebesgue measurable subsets of [a, b], and µ is the usual Lebesgue
measure on A. Set
n Z o
M := [g] ∈ H : g ∈ L2 , g dµ = 0 .
[a,b]
70
4.6. Exercises
so H = H
f +̇ H
1
f (cf. Exercise 3.5).
2
Check that (H
f )⊥ = H
1
f and (H
2
f )⊥ = H
2
f . Deduce that the projection
1
of H on H
f along H
1
f is the orthogonal projection of H on H
2
f.
1
(a) H is separable;
Note that Example 4.2.5 shows that (b) ⇒ (c). So it suffices to show that
(a) ⇒ (b), and (c) ⇒ (a).
Exercise 4.10. Let H1 and H2 be Hilbert spaces over F, and let H be the
(external) direct product of H1 and H2 , as defined in Exercise 4.7. Assume
B1 and B2 are orthonormal bases for H1 and H2 , respectively.
Find an orthonormal basis B for H in terms of B1 and B2 .
71
4. More on Hilbert spaces
Exercise 4.11. Let H = L2 ([−1, 1], A, µ), where A denote the Lebesgue-
measurable subsets of [−1, 1] and µ is the restriction of the usual Lebesgue
measure to A.
For each nn ∈ {0} ∪ N, let pn+1 o: [−1, 1] → C be defined by pn+1 (t) = tn ,
and set S := [pn+1 ] : n ∈ {0} ∪ N ⊆ H.
a) Show that Span (S) is dense in H.
b) Apply the Gram-Schmidt
n orthonormalization
o process to S to obtain
an orthonormal basis B = [qn+1 ] : n ∈ {0} ∪ N for H, where each qn+1 is
the polynomial on [−1, 1] given by
q
n+ 1
2 dn 2
qn+1 (t) = (t − 1)n
.
2n n! dtn
(These polynomials are called the normalized Legendre polynomials.)
converges to x when the following holds: given ε > 0, there exists a finite
subset F0 ⊆ J such that for all finite subsets F of J containing F0 , we have
X
x− xj k < ε ,
j∈F
x=
X
hx, ui u .
u∈B
PM (x) =
X
hx, vi v .
v∈C
72
4.6. Exercises
n∈Z
c) Set g(t) = et for all t ∈ [−π, π]. Use Parseval’s identity to obtain a
formula for the sum of the series
∞
X 1
.
n=1 n2 +1
We recall that f : [−π, π] → C is called even if f (−t) = f (t) for all t, while
it is called odd if f (−t) = −f (t) for all t.
a) Show that Heven is a closed subspace of H and that (Heven )⊥ = Hodd .
Describe the orthogonal projection P of H on Heven .
Hint: It might be helpful to consider the map [f ] → [fe ], where fe(t) := f (−t).
b) Find an orthonormal basis for Heven and one for Hodd .
73
4. More on Hilbert spaces
Exercise 4.16. Show that the formula for (TK )∗ in Example 4.3.7 is correct.
Hint : Consider H0 = [g] : g ∈ C([a, b])} and use Exercise 4.15.
Exercise 4.17. Prove Proposition 4.3.10.
Exercise 4.18. Let v, w ∈ H and consider the linear operator Tv,w : H → H
defined by
Tv,w (x) := hx, vi w for all x ∈ H .
Note that Tv,w has rank one if v, w ∈ H \ {0}.
a) Show that Tv,w is bounded with norm kTv,w k = kvk kwk. Then show
that (Tv,w )∗ = Tw,v .
b) Show that every T ∈ B(H) which has rank one is of the form T = Tv,w
for some v, w ∈ H \ {0}.
c) Assume T ∈ B(H) is a finite-rank operator, T 6= 0. Show that T may
be written as a finite sum of rank one operators in B(H).
Hint : Start by picking an orthonormal basis for T (H).
d) Show that if T ∈ B(H) is a finite-rank operator, then so is T ∗ .
Exercise 4.19. Let T ∈ B(H) and let M be a closed subspace of H.
Show that
M is invariant under T if and only if M ⊥ is invariant under T ∗ .
74
4.6. Exercises
Exercise 4.21. Let H and K be Hilbert spaces over F, and let T ∈ B(H, K).
a) Show that there exists a unique operator T ∗ ∈ B(K, H) (called the
adjoint of T ) satisfying that
D E D E
T (x), y = x, T ∗ (y) for all x ∈ H and all y ∈ K.
D c) Assume
E F = C and let T ∈ B(H). Show that T = 0 if and only if
T (x), x = 0 for all x ∈ H.
75
4. More on Hilbert spaces
(i) T is self-adjoint;
(ii) WT ⊆ R;
D E
(iii) T (x), x ∈ R for all x ∈ H.
kSk ≤ 1 ⇔ (IH − S ∗ S) ≥ 0 .
Exercise 4.28. Let H = L2 ([0, 1]) (with usual Lebesgue measure) and let
T = Mf be the self-adjoint operator in B(H) given by multiplication with
the function f (t) = t on [0, 1], cf. Example 4.4.4. Show that T has no
(complex) eigenvalues.
76
4.6. Exercises
Exercise 4.30. Assume H 6= {0} is separable (cf. Exercise 4.9) and infinite-
dimensional. Let then B be an orthonormal basis for H indexed by Z, say
B = {vk }k∈Z . One may then define the bilateral shift operator V : H → H
(w.r.t. B) by
n
V (x) = n→∞
lim for all x ∈ H.
X
hx, vk i vk+1
k=−n
77
CHAPTER 5
On compact operators
for all s ∈ [0, 1]. It follows that T (g) ∈ X. Moreover, the map T : X → X
sending g to T (g) is clearly linear. As T (X) is 2-dimensional, T has finite-
rank. Further, since
Z 1 Z 1
[T (g)](s) ≤ | sin(s − t)g(t)| dt ≤ |g(t)| dt ≤ kgku
0 0
for all s ∈ [0, 1], we get that kT (g)ku ≤ kgku for all g ∈ X. Hence, T is
bounded. We can therefore conclude that T is compact.
80
5.1. Introduction to compact operators between normed spaces
81
5. On compact operators
kT (x0l )−T (x0k )k ≤ kTm (x0l )−Tm (x0k )k+2M kT −Tm k < ε/3+2M (ε/3M ) = ε
for all k, l ≥ N . Hence we have shown that the claim (5.1.1) is true.
Finally, as Y is a Banach space, we know that B(X, Y ) is a Banach
space too, and this implies that K(X, Y ), being closed in B(X, Y ), is also a
Banach space.
Then we have
F(X, Y ) ⊆ K(X, Y ).
for all x ∈ X and all n ∈ N. One readily checks that kMλ k = kλk∞ .
Now, assume that λ ∈ c0 (N), i.e., limn→∞ λ(n) = 0. Then Mλ is compact.
Indeed, for each k ∈ N, let λ(k) ∈ `∞ (N) be defined by
λ(n) if 1 ≤ k ≤ n,
λ(k) (n) =
0 otherwise,
for every n ∈ N. Then it is clear that each Mλ(k) has finite-rank; moreover,
82
5.2. On compact operators on Hilbert spaces
This property implies that no operator in K(X) can have a bounded inverse
when X is infinite-dimensional (for if T ∈ K(X) has an inverse T −1 ∈ B(X),
then we must have that IX = T −1 T ∈ K(X), so dim(X) < ∞).
We end this section with an interesting result concerning the possible
eigenvalues of a compact operator.
F(X, H) = K(X, H) .
83
5. On compact operators
Proof. By Corollary 5.1.6, we only have to show that K(X, H) ⊆ F(X, H).
So let T ∈ K(X, H), and let ε > 0. We need to prove that there exists
S ∈ F(X, H) such that kT − Sk ≤ ε. Clearly, we can assume T 6= 0.
Set A := T (X1 ). Since X1 is bounded and T is compact, the set A
is compact in H. As H is a metric space, this implies that A is totally
bounded (cf. Proposition 3.5.12 in Lindstrøm’s book). Hence we can cover
A with some open balls B1 , . . . , Bn of radius ε/4, having respective centers
a1 , . . . , an ∈ A. For each j = 1, . . . , n, we can then find xj ∈ X1 such that
kaj − T (xj )k < ε/4.
Set now F := Span ({T (x1 ), . . . , T (xn )}), which is a finite dimensional
subspace of H, and let PF denote the orthogonal projection of H on F . Since
the range of PF T is contained in F , PF T has finite-rank, so PF T ∈ F(X, H).
We claim that
kT − PF T k ≤ ε .
Indeed, let x ∈ X1 . Then T (x) ∈ A, so T (x) ∈ Bj for some j ∈ {1, . . . , n}.
Hence,
F(H) = K(H) .
84
5.2. On compact operators on Hilbert spaces
j=1 j=1
k=1
85
5. On compact operators
Note that the change of order of summation at the second but last step
above is allowed since we are dealing with sums of non-negative numbers.
Applying what we have done to the case where B = C, i.e., uj = vj for every
j ∈ N, we get that
∞ ∞
kT (vj )k2 = kT ∗ (vk )k2 .
X X
j=1 k=1
as desired.
j=1
kT k ≤ kT k2
86
5.2. On compact operators on Hilbert spaces
Proof. We first note that it is evident from the proof of Lemma 5.2.5 that
T ∗ ∈ HS(H) whenever T ∈ HS(H).
Let B = {uj }j∈N be an orthonormal basis for H, and let T, T 0 ∈ HS(H).
Define ξ, ξ 0 ∈ `2 (N) by
ξ(j) := kT (uj )k and ξ 0 (j) := kT 0 (uj )k for each j ∈ N,
so that kξk2 = kT k2 and kξ 0 k2 = kT 0 k2 . Using the triangle inequality, first
in H, and then in `2 (N), we get
∞ ∞ 2
k(T + T )(uj )k ≤
0 2
kT (uj )k + kT 0 (uj )k = kξ + ξ 0 k22
X X
j=1 j=1
Thus, kT k ≤ kT k2 . j=1
j=1 j=1
Since Tn ∈ F(H) for each n, Theorem 5.1.6 gives that T ∈ K(H). Hence,
HS(H) ⊆ K(H).
It only remains to show that F(H) ⊆ HS(H), but we leave this as an
exercise.
87
5. On compact operators
(which does not depend on the choice of orthonormal basis for H).
Letting A = [ai,j ] denotes the matrix of T w.r.t. B, one readily checks
that
n
X 1/2
kT k2 = |ai,j |2 ,
i,j=1
Example 5.2.11. Set H = L2 ([a, b], A, µ), where A denotes the Lebesgue
measurable subsets of a closed interval [a, b] and µ is the Lebesgue measure
on A. Let K : [a, b] × [a, b] → C be a continuous function and let TK ∈ B(H)
denote the associated integral operator on H, which is the extension of the
integral operator TK : C([a, b]) → C([a, b]) given by
Z b
[TK (f )](s) = K(s, t) dt for f ∈ C([a, b]) and s ∈ [a, b].
a
88
5.2. On compact operators on Hilbert spaces
j=1 j=1
89
5. On compact operators
λ = λ hx, xi = hλ x, xi = hT (x), xi ∈ WT ⊆ R ,
90
5.3. The spectral theorem for a compact self-adjoint operator
Hence,
T (y) = lim T (λxn ) = λ lim T (xn ) = λ y .
n→∞ n→∞
We are now ready for the spectral theorem for a compact self-adjoint
operator T . Intuitively, we could hope to be able to construct an orthonormal
basis of eigenvectors for T by using Lemma 5.3.3 repeatedly as follows. Start
by picking a unit eigenvector v0 of T associated to the eigenvalue λ0 = ±kT k.
Next, consider the restriction T1 of T to {v0 }⊥ , and pick a unit eigenvector
v1 of T1 associated to the eigenvalue λ1 = ±kT1 k. Then continue this process
inductively. There are several technicalities involved in working out the
details of this approach. We will follow a more pedestrian route, which also
provides more information about T .
91
5. On compact operators
– T = λ Pλ if L is finite ;
P
λ∈L
92
5.3. The spectral theorem for a compact self-adjoint operator
93
5. On compact operators
x = x M + xM ⊥ ,
lim (xn + yn ) = xM + xM ⊥ = x .
n→∞
E0 = Eλ0 k = {vk,l : k ∈ N, 1 ≤ l ≤ nk }.
[
k∈N
Let now ε > 0. We have to show that there exists N ∈ N such that
kT − nk=1 λk Pλk k ≤ ε for all n ≥ N .
P
94
5.3. The spectral theorem for a compact self-adjoint operator
Using (b), we can choose N ∈ N such that |λk | < ε for all k > N .
Then for all n ≥ N and all x ∈ H, using continuity of the norm in H and
Pythagoras’ identity, we get
n ∞ ∞
2 2
λk Pλk (x) = λk Pλk (x) = |λk |2 kPλk (x)k2
X X X
T−
k=1 k=n+1 k=n+1
∞
2
kPλk (x)k2 ≤ ε2 kxk2
X
≤ε
k=n+1
for every j ∈ N . Since each µj is nonzero, we get that all µj ’s are positive.
For each j ∈ N , set
√ 1
σj := µj > 0 and uj := S(vj ) .
σj
95
5. On compact operators
j∈N
so
S(x) = hx, vj i S(vj ) + S(z) =
X X
σj hx, vj i uj ,
j∈N j∈N
as asserted in (5.3.3).
It readily follows that {uj : j ∈ N } is an orthonormal basis for S(H).
√
Finally we remark that the spectral theorem also gives that σj = µj → 0
as j → ∞ when N is countably infinite, and that
kSk = kT k1/2 = max{ µj : j ∈ N }1/2 = max{ σj : j ∈ N } .
96
5.4. Application: The Fredholm Alternative
which is easily verified for any subspace M of H, and the dimension formula
for F . We get that
97
5. On compact operators
Proof. Assume first that µ is not an eigenvalue of T , i.e., ker(T − µI) = {0}.
Then the spectral theorem implies that the equation (T − µI)(x) = y
has a unique solution for all y ∈ H. (You are asked to check this in
Exercise 5.9.) Thus, F = T − µI is surjective, and this implies that
ker(F ∗ ) = ker(T − µI) = {0}, i.e., µ is not an eigenvalue of T . Arguing as
above, we get that the equation (T − µI)(x0 ) = y 0 , i.e., (T − µI)∗ (x0 ) = y 0
has a unique solution for all y 0 ∈ H. This shows that (a) in Definition 5.4.1
holds in this case.
Next, assume that µ is an eigenvalue of T , i.e., ker(T − µI) 6= {0}. Then
µ ∈ R, so F ∗ = F . Moreover, as µ 6= 0, we have that T 6= 0, and the spectral
theorem tells us that 1 ≤ dim(ker(F )) = dim(ker(T − µI)) < ∞. Hence,
to show that (b) in Definition 5.4.1 holds, it remains only to prove that the
1
Such equations, and Fredholm integral equations of the first kind (i.e., equations of
the form TK (f ) = g), were studied by I. Fredholm at the beginning of the 20th century.
They arise in some practical problems in signal theory and in physics.
98
5.4. Application: The Fredholm Alternative
j∈J
99
5. On compact operators
5.5 Exercises
Exercise 5.1. Let X, Y, Z denote normed spaces over F. Consider λ ∈ F,
T, T 0 ∈ B(X, Y ) and S ∈ B(Y, Z), so ST ∈ B(X, Z).
a) Show that λ T + T 0 ∈ K(X, Y ) if T, T 0 ∈ K(X, Y ).
b) Show that ST ∈ K(X, Z) if T ∈ K(X, Y ).
c) Show that ST ∈ K(X, Z) if S ∈ K(Y, Z).
d) Set K(X) = K(X, X). Deduce that
ST ∈ K(X) if S ∈ B(X) and T ∈ K(X), or if S ∈ K(X) and T ∈ B(X).
Exercise 5.2. Let X = `p (N), λ ∈ `∞ (N), and Mλ ∈ B(X) be the
associated multiplication operator, cf. Example 5.1.7.
Show that λ ∈ c0 (N) if Mλ is compact.
(It therefore follows that Mλ is compact if and only if λ ∈ c0 (N).)
Exercise 5.3. Let X be a normed space, H be a Hilbert space, and let
T ∈ K(X, H). Show that T (X) is separable.
Exercise 5.4. Let H beD an infinite-dimensional
E Hilbert space and let
T ∈ K(H). Show that T (un ), un → 0 as n → ∞ whenever {un }n∈N
is an orthonormal sequence in H.
Exercise 5.5. Let H be a Hilbert space and let P ∈ B(H) be a projection
(i.e. P 2 = P ). Show that P has finite-rank if (and only if) P is compact.
Exercise 5.6. Let H be a separable Hilbert space, H 6= {0}.
a) Show that F(H) ⊆ HS(H), and that F(H) is dense in HS(H)
w.r.t. k · k2 .
b) Assume that T ∈ HS(H) and S ∈ B(H). Show that both ST and
T S belong to HS(H), and that we have
Show that this gives a well-defined inner product on HS(H), and check that
the associated norm is the Hilbert-Schmidt norm k · k2 .
100
5.5. Exercises
X hy, uj i
uj
j∈J µj − µ
converges to some h ∈ H.
b) Set z := y − PM (y) and x := h − 1
µ
z. Show that (T − µIH )(x) = y.
101
5. On compact operators
102
5.5. Exercises
L := {λ ∈ R | λ is a nonzero eigenvalue of T }.
Set also
e := {λ ∈ R | λ is an eigenvalue of T },
L
so L = L
e \ {0}. Show the following assertions:
(i) L
e = {µ | j ∈ N} and L = {µ | j ∈ N, µ 6= 0}.
j j j
Exercise 5.17. Let H = L2 ([0, 1]) (with usual Lebesgue measure) and let
T = Mf be the self-adjoint operator in B(H) given by multiplication with
the function f (t) = t on [0, 1], cf. Example 4.4.4.
Show that T (H) is not closed, i.e., that T does not have closed range.
Show also that T is not compact.
103
5. On compact operators
Exercise 5.19. Let H be a Hilbert space and T ∈ B(H). Let us say that T
is bounded from below if there exists some α > 0 such that α kxk ≤ kT (x)k
for all x ∈ H. For example, T is bounded from below when T is an isometry.
Show that if T is bounded from below, then T has closed range.
Exercise 5.20. Finish the proof of Theorem 5.9 by checking that the
sequence {cj }j∈J constructed in the final paragraph (under the assumption
that y is orthogonal to Eµ ) belongs to `2 (J).
104