0% found this document useful (0 votes)
4 views12 pages

NCG 3

Chapter 3 discusses operators on a Hilbert space, focusing on positive operators and their properties. It presents the polar decomposition theorem, which states that any operator can be expressed as the product of a partial isometry and a positive operator. The chapter also includes various propositions and corollaries that establish relationships between self-adjointness, positivity, and the spectra of operators.

Uploaded by

Angelo Oppio
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
4 views12 pages

NCG 3

Chapter 3 discusses operators on a Hilbert space, focusing on positive operators and their properties. It presents the polar decomposition theorem, which states that any operator can be expressed as the product of a partial isometry and a positive operator. The chapter also includes various propositions and corollaries that establish relationships between self-adjointness, positivity, and the spectra of operators.

Uploaded by

Angelo Oppio
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 12

CHAPTER 3

Operators on a Hilbert Space

This chapter is a review of basic results concerning operators on a Hilbert space.


The main reference for this chapter is the book of Reed-Simon [RS].
Throughout this chapter we let H be a separable Hilbert space and we denote
by L(H) its C ∗ -algebra of continuous endomorphisms.

3.1. Polar Decomposition


Definition 3.1.1. An operator T ∈ L(H) is said to be positive if can be written
in the form T = S ∗ S with S ∈ L(H).
We denote by L(H)+ the set of positive elements of L(H).
Proposition 3.1.2. For T ∈ L(H) the following are equivalent:
(i) T is positive.
(ii) There exists S ∈ L(H) selfadjoint such that T = S 2 .
(iii) T is selfadjoint and Sp T ⊂ [0, ∞).
(iv) hT ξ, ξi ≥ 0 for all ξ ∈ H.
Proof. We shall prove the implications (i) ⇒ (iv), (ii) ⇒ (i), (iii) ⇒ (ii) and
(iv) ⇒ (iii), which will give the proposition.
• (i) ⇒ (iv): If T = S ∗ S for some S ∈ L(H) then, for all ξ ∈ H,

hT ξ, ξi = hS ∗ Sξ, ξi = hSξ, Sξi = kSξk2 ≥ 0,


proving that (i) ⇒ (iv).
• (ii) ⇒ (i): If T = S 2 with S ∈ L(H) selfadjoint, then T = S ∗ S and hence T is
positive. Thus (ii) ⇒ (i).
• (iii) ⇒ (ii): Assume that T is selfadjoint and Sp T ⊂ [0, ∞). Then the function
√ √
t is defined and continuous on Sp T ,√and so we can define S := T by functional
continuous calculus. Notice that, as t takes real values on Sp T , it follows from
Remark 1.5.4 that S is selfdajoint. Recall that the continuous functional √ 2calculus
f → f (T ) is a homomorphism √ of algebrs from C(Sp T ) to L(H). Since ( t) = t on
Sp T , it follows that S 2 = ( T )2 = T , and hence T satisfies (ii). Thus (iii) ⇒ (ii).
• (iv) ⇒ (iii): Assume that hT ξ, ξi ≥ 0 for all ξ ∈ H. Then T is selfadjoint. Indeed,
for all ξ, η in H, we have
4hT ξ, ηi = hT (ξ +η), ξ +ηi−hT (ξ −η), ξ −ηi−ihT (ξ +iη), ξ +iηi+ihT (ξ −iη), ξ −iηi,
from which we see that hT ξ, ηi = hT η, ξi, and hence hT ξ, ηi = hξ, T ηi.
Let us now show that Sp T ⊂ [0, ∞). Since T is selfadjoint, and hence Sp T ⊂ R
by Proposition 1.3.4, we only need to show that T − λ is invertible for all λ < 0.
1
Thus let λ ∈ (−∞, 0) and let ξ ∈ H. As hT ξ, ξi ≥ 0 we have
|λ|kξk2 = −λhξ, ξi ≤ h(T − λ)ξ, ξi ≤ k(T − λ)ξkkξk.
Thus,
(3.1) k(T − λ)ξk ≤ |λ|kξk ∀ξ ∈ H.
This implies that ker(T − λ) = {0}. As T − λ is selfadjoint we then see that
im(T − λ) = (ker(T − λ))⊥ = H.
In fact, the inequality (3.1) also implies that im(T − λ) is closed. Indeed, if
η = lim(T − λ)ξn , then (3.1) implies that the sequence (ξn )n≥0 is Cauchy in H, and
hence ξn converges to some ξ in H. Then η = lim(T − λ)ξn = T ξ, showing that η
is contained in im(T − λ). Thus ker(T − λ) = 0 and im(T − λ) = H, i.e., T − λ is
bijective. Recall that by the open mapping theorem any bijective continuous linear
map of H onto H has a continuous inverse (see [Fo, p. 162]), so T −λ is an invertible
element of L(H). Thus Sp T ⊂ [0, ∞), and hence T satisfies (iii), showing that (iv)
implies (iii). The proof is complete. 
Corollary 3.1.3. The set of positive operators L(H)+ is a positive cone of
L(H), i.e.,
λ1 T1 + λ2 T2 ∈ L(H)+ ∀Tj ∈ L(H)+ ∀λj ≥ 0.
Proof. For j = 1, 2 let Tj ∈ L(H)+ and let λj ∈ [0, ∞). Using the character-
ization (iv) of Proposition 3.1.2, we see that, for all ξ ∈ H,
h(λ1 T1 + λ2 T2 )ξ, ξi = λ1 hT1 ξ, ξi + λ2 hT2 ξ, ξi ≥ 0,
proving that λ1 T1 + λ2 T2 is a positive operator. 
Corollary 3.1.4. Let T ∈ L(H) be normal and let f ∈ C(Sp T ) be non-
negative. Then the operator f (T ) is positive.
Proof. Since f is real-valued it follows from Remark 1.5.4 that f (T ) is selfad-
joint. Moreover, by (1.13) we have Sp f (T ) = f (Sp T ) ⊂ [0, ∞), so it follows from
Proposition 3.1.2 (iii) that f (T ) is positive. 
Let T ∈ L(H). Then T ∗ T is a positive operator, so by the previous proposition

T T is selfadjoint and its spectrum is contained in√[0, ∞). Therefore, by continuous
functional calculus we can define its square root T ∗ T as an element of L(H). It
follows from Corollary 3.1.4 that |T | is a positive operator.

Definition 3.1.5. For all T ∈ L(H) the operator T ∗ T is denoted |T | and is
called the modulus of T .
Lemma 3.1.6. Let T ∈ L(H). Then
(i) |T | is the unique element of L(H)+ whose square is T ∗ T .
(ii) We have
(3.2) k|T |ξk = kT ξk ∀ξ ∈ H,
and hence ker |T | = ker T .
Proof. The continuous functional
√ 2 calculus f → f (T ∗ T ) is a ∗-homomorphism


from C(Sp T T ) to L(H). As ( t) = t on [0, ∞) it follows that |T |2 = ( T ∗ T )2 =
T ∗T .
2
Let S ∈ L(H)+ be such that S 2 = T . Since S√is positive, by Proposition 3.1.2
its spectrum is contained in [0, ∞),√and hence √ t2 = t on Sp S. Therefore, by
2 ∗
continuous functional calculus S = S = T T = |T |. Thus |T | is the unique
element of L(H)+ whose square is T ∗ T .
As |T | is selfadjoint |T |∗ |T | = |T |2 = T ∗ T . Therefore, for all ξ ∈ H,
k|T |ξk2 = h|T |ξ, |T ||ξ|i = h|T |∗ |T |ξ, ξi = hT ∗ T ξ, ξi = hT ξ, T ξi = kT ξk2 ,
proving (3.2). This immediately implies that |T | and T have same kernel. 
Proposition 3.1.7 (Polar Decomposition). Let T ∈ L(H). Then there exists
a unique U ∈ L(H), called the phase of T , such that
(i) T = U |T |;
(ii) ker U = ker |T |.
Proof. Since |T is selfadjoint (ker |T |)⊥ = im |T |, and hence |T | is a bijection
from im |T | onto itself. Let |T |−1 : im |T | → im |T | be its inverse and denote by U
the linear map T |T |−1 : im |T | → H. Then, for any ξ ∈ im |T |, we have
kU ξk2 = hT |T |−1 ξ, T |T |−1 ξi = hT ∗ T |T |−1 ξ, |T |−1 ξi = h|T |2 |T |−1 ξ, |T |−1 ξi
= h|T |ξ, |T |−1 ξi = hξ, |T ||T |−1 ξi = kξk2 .
Thus U uniquely extends to an isometric linear map U : im |T | → H. Extending U
to be 0 on ker |T | = (im |T |)⊥ we then get a continuous endomorphism U : H → H
whose null space is ker |T |.
If ξ ∈ im |T |, then U |T |ξ = T |T |−1 |T |ξ = ξ, and hence T = U |T | on im |T | by
continuity. Moreover, as by Lemma 3.1.6 ker |T | = ker T , we have U |T |ξ = 0 = T ξ
for all ξ in ker |T | = (im |T |)⊥ . Therefore, we see that T = U |T | on H, showing
that U satisfies the conditions (i) and (ii) of the proposition.
Let V ∈ L(H) be such that T = V |T | and ker V = ker |T |. If ξ ∈ ker |T |, then
obviously V ξ = 0 = U ξ. If ξ ∈ im |T |, then V ξ = V |T ||T |−1 ξ = T |T |−1 ξ = U ξ, so
by continuity V = U on im |T |. It follows from this that V = U on H, and hence
U is the unique element of L(H) such that T = U |T | and ker U = ker |T |, giving
the proposition. 
Proposition 3.1.8. Let T ∈ L(H) have polar decomposition T = U |T | and
denote by Π0 (T ) (resp. Π0 (T ∗ )) the orthogonal projection onto ker T (resp. ker T ∗ ).
(i) The range of U is im T .
(ii) We have
U ∗ U = 1 − Π0 (T ) and U U ∗ = 1 − Π0 (T ∗ ),
so that U is a partial isometry and has norm 1 unless T = 0.
(iii) If T is injective and has dense range, then U is unitary.
(iv) We have
|T | = U ∗ T, T ∗ = U T U ∗, |T ∗ | = T U ∗ = U |T |U ∗ .
(v) The phase of T ∗ is U ∗ .
Proof. As U vanishes on ker |T | we see that im U = U (ker |T |). Notice that
by its construction in the proof of Proposition 3.1.7 the operator U is isometric
on (ker |T |)⊥ and agrees with T |T |−1 on im |T |. In particular, it follows from
Lemma 1.1.8 that U ((ker |T |)⊥ ) is closed and U induces a unitary operator from
3
(ker |T |)⊥ onto U ((ker |T |)⊥ ) = im U . As (ker |T |)⊥ = im T we then see that
im U = U (im |T |). Since U = T |T |−1 on im |T | we have U (im |T |) ⊂ im T , and
hence im U ⊂ im T . However, as T = U |T | we also have im T ⊂ im U , and as im U
is closed we see that im T is contained is im U , and so the range of U is im T .
As abovementioned U induces a unitary operator from (ker |T |)⊥ onto im U =
im T . Since by Lemma 3.1.6 ker |T | = ker T this shows that U is a unitary endo-
morphism of H when T is injective and has dense range. In any case, as we have
im T = (ker T ∗ )⊥ we see that U induces a unitary operator from (ker T )⊥ onto
(ker T ∗ )⊥ . Since ker U = ker |T | = ker T we then deduce that U ∗ U is the orthogo-
nal projection onto (ker T )⊥ and U U ∗ is the orthogonal projection onto (ker T )⊥ ,
that is, U ∗ U = 1 − Π0 (T ) and U U ∗ = 1 − Π0 (T ∗ ). In particular, if T 6= 0 then
kU k2 = kU ∗ U k = k1 − Π0 (T )k = 1, i.e., kU k = 1.
Notice that, as ker |T | = ker T , we have |T |(1 − Π0 (T )) = (1 − Π0 (T ))|T | = |T |,
and hence U ∗ T = U ∗ U |T | = (1 − Π0 (T ))|T | = |T |. Moreover,
(U |T |U ∗ )2 = U |T |U ∗ U |T |U ∗ = U |T |(1 − Π0 (T ))|T |U ∗ = (U |T |)(U |T |)∗ = T T ∗ .
As |T | is positive, for any ξ ∈ H, we have hU |T |U ∗ ξ, ξi = h|T |U ∗ ξ, U ∗ ξi ≥ 0, and
hence U |T |U ∗ is positive. Thus U |T |U ∗ is a positive operator whose square is equal
to T T ∗ , so using Lemma 3.1.6 we see that U |T |U ∗ = |T ∗ |. Since T = U |T | this
also shows that |T ∗ | = T U ∗ .
Notice that ker U ∗ = (im U )⊥ = (im T )⊥ = ker T ∗ = ker |T ∗ |. Moreover,
U ∗ |T ∗ | = U ∗ U |T |U ∗ = (1 − Π0 (T ))|T |U ∗ = |T |U ∗ = (U |T |)∗ = T ∗ ,
so by Proposition 3.1.7 the phase of T ∗ is U ∗ . Notice that the above equalities in-
clude the equality T ∗ = |T |U ∗ . As |T | = U ∗ T we see that T ∗ = U ∗ T U ∗ , completing
the proof. 

3.2. Spectral Theorem and Borel Functional Calculus


Let T ∈ L(H) be a normal operator (i.e., T ∗ T = T T ∗ ) and set S = Sp T .
Theorem 3.2.1 (Spectral Theorem; see [RS, Thm. VII.3]). There exist a fi-
nite measure space (X, µ), a unitary operator U : H → L2µ (X), and a bounded
measurable function f on X, in such way that
(3.3) U T U ∗ξ = f ξ ∀ξ ∈ L2µ (X).
For F ∈ L∞ (X) denote by TF the multiplication operator by F , i.e., the
operator TF ∈ L(L2µ (X)) defined by
TF ξ = F ξ ξ ∈ L2µ (X).
For instance, Eq. (3.3) says that U T U ∗ = Tf .
The essential range of F consists of all λ ∈ C such that
µ(λ −  < F < λ + ) > 0 ∀ > 0.
It can be shown that Sp TF agrees with the essential range of F . In particular, we
see that the essential range of Tf is S. Thus, without any loss of generality, we may
assume that the range of f is S.
We endow L∞ (X) with its usual norm, i.e.,
kF kL∞ (X) = {λ; λ in the essential range of |F |} ∀F ∈ L∞ (R).
4
We also endow L∞ (X) with the involution F → F given by complex conjugation.
This turns L∞ (X) into a commutative unital C ∗ -algebra. Then it is not difficult
to check that the map F → TF is a ∗-homomorphism from L∞ (X) to L(L2µ (X)).
Moreover, as TF is a normal operator, we have
kTF k = sup |λ| = kF kL∞ (S) .
λ∈Sp TF

If g ∈ C(S) then g ◦ f is again a bounded measurable function on X. In fact,


the continuous functional calculus for f is just g → g ◦ f . Since the maps F → TF
(resp., F → U ∗ TF U ) is an isometric ∗-homomorphisms from L∞ (X) to L(L2µ (X))
(resp., L(H)), it follows that, for all g ∈ C(S),
(3.4) g(T ) = g(U ∗ Tf U ) = U ∗ g(TF )U = U ∗ Tg◦f U.
We then can extend the definition of g(T ) for any any bounded Borel function g
on Sp T by letting
g(T ) := U ∗ Tg◦f U.
This defines a bounded operator on H.
Theorem 3.2.2 (Borel Functional Calculus; see [RS, Thm. VII.2]). The fol-
lowing holds.
(1) The map g → g(T ) is a ∗-homomorphism from L∞ (S) to L(H) such that
(3.5) kg(T )k ≤ kgkL∞ (S) ∀g ∈ L∞ (S).
(2) If (gn )n≥0 is a bounded sequence in L∞ (S) such that gn → g a.e., then
gn (T ) → g(T ) strongly (i.e., gn (T )ξ → g(T )ξ for all ξ ∈ H).
(3) If g ∈ L∞ (S) is real-valued (resp., non-negative), then g(T ) is selfadjoint
(resp., positive).

3.3. Unbounded Operators


The spectral theorem and the Borel functional calculus can be extended to
unbounded operators as follows.
Definition 3.3.1. An (unbounded) operator on H is a linear operator T :
D(T ) → H, where the domain D(T ) is a subspace of H. T
An operator T on H is said to be densily defined when its domain D(T ) is a
dense subspace of H. An operator S is said to be an extension of T , and we write
T ⊂ S, if D(T ) ⊂ D(S) and S agrees with T on D(T ).
The graph of an operator T is defined to be
G(T ) = {(ξ, η) ∈ H ⊕ H; η = T ξ}.
The graph of T is a subspace of H ⊕ H. We say that T is closed when G(T ) is a
closed subspace of H ⊕ H. We say that T ∗ is closable when G(T ) is the graph of
an operator T . In this case we call T the closure of T . This is the smallest closed
extension of T .
If T is densily defined, then its adjoint is the operator T ∗ with graph
G(T ∗ ) = {(ξ, η) ∈ H ⊕ H; hT ζ, ξi = hζ, ηi ∀ζ ∈ D(T )}.
Since G(T ∗ ) is a closed subspace, we see that T ∗ is always a closed operator.
The resolvent set of a closed operator T consists of all λ ∈ C such that
(i) T − λ : D(T ) → H is a bijection.
5
(ii) The inverse (T − λ)−1 : H → D(T ) is bounded.
The spectrum of T , denoted Sp T , is the complement of the resolvent set.
It can be shown that Sp T is a closed subset of C (which may be empty) and
the resolvent λ → (T − λ)−1 is analytic from C \ Sp T to L(H) (provided we regard
the inverses (T − λ)−1 as elements of L(H)).
A densily defined operator T is said to be symmetric is T ⊂ T ∗ . It is said to
be selfadjoint if T = T ∗ . Note that any symmetric operator is closable. We also
say that T is essentially selfadjoint when T is symmetric and closable.
Proposition 3.3.2 ([RS, Thm. VIII.3]). Let T be a symmetric operator on H.
Then the following are equivalent:
(i) T is selfadjoint.
(ii) T is closed and ker(T ± i) = {0}.
(iii) im(T ± i) = H.
Corollary 3.3.3 (see [RS]). Let T be a symmetric operator on H. Then the
following are equivalent:
(i) T is essentially selfadjoint.
(ii) ker(T ± i) = {0}.
(iii) im(T ± i) = H.
Let T be a selfadjoint (unbounded) operator on H.
Theorem 3.3.4 (Spectral Theorem; see [RS, Thm. VIII.4]). There exist a
measured space (X, µ) with µ(X) < ∞, a unitary operator U : H → L2µ (X), and a
measurable real-valued function f on X such that
U (D(T )) = {ξ ∈ L2µ (X); f ξ ∈ L2µ (X)},
U T U ∗ξ = f ξ ∀ξ ∈ U (D(T )) .
If g is a bounded Borel function on R, then we define g(T ) as the bounded
operator on L(H) given by
g(T ) := U ∗ Tg◦f U.
Theorem 3.3.5 (Borel Functional Calculus; [RS, Thm. VIII.5]). The following
holds.
(1) The map g → g(T ) is a ∗-homomorphism from L∞ (R) to L(H) such that
(3.6) kg(T )k ≤ kgkL∞ (R) ∀g ∈ L∞ (R).
(2) If (gn )n≥0 is a bounded sequence in L∞ (R) such that gn → g a.e., then
gn (T ) → g(T ) strongly.
(3) If g ∈ L∞ (R) is real-valued (resp., non-negative), then g(T ) is selfadjoint
(resp., positive).
More generally, if g is a possibly unbounded Borel function on R, then g(T )
makes sense as an unbounded operator as follows. The domain of g(T ) is
 
∗ 2 2
D(g(T )) = U {ξ ∈ Lµ (X); (g ◦ f )ξ ∈ Lµ (X)} ,

and we define g(T ) by the formula


g(T )ξ := U ∗ ((g ◦ f )U ξ) ∀ξ ∈ D(g(T )).
6
For instance, if g(t) = t, then g(T ) = T . In addition, it is not hard to check that,
if (gn ) be a sequence of Borel functions on R such that gn → g and |gn | ≤ |g|, then
as n → ∞ we have
gn (T )ξ −→ g(T )ξ ∀ξ ∈ D(g(T )).
Example 3.3.6. Let ∆ = −(∂x21 + . . . + ∂x2n ) be the (positive) Laplacian on Rn .
We shall regard ∆ as an operator on the space S 0 (Rn ) of tempered distributions on
Rn . Denoting by u → û the Fourier transform on S 0 (Rn ), we have
(3.7) (∆u)∧ = |ξ|2 û ∀u ∈ S 0 (Rn ).
Thus, we can regard ∆ as an unbounded operator on the Hilbert space H = L2 (Rn )
with domain,
D(∆) = {u ∈ L2 (Rn ); |ξ|2 û ∈ L2 (Rn )}.
Notice that D(∆) agrees with the Sobolev space W 2,2 (Rn ). As such ∆ is a selfad-
joint operator with spectrum [0, ∞).
Denote by U the unitary operator on L2 (Rn ) defined by
n
U v = (2π)− 2 v̂ ∀v ∈ L2 (Rn ).
Then it follows from (3.7) that
∆v = U ∗ T|ξ|2 U v ∀v ∈ W 2,2 (Rn ).
Therefore, we see that in this example the spectral theorem follows from elementary
Fourier-analytic considerations.
If g is a bounded Borel function on R, then g(∆) is given by
Z
g(∆)v(x) = (U ∗ Tg(|ξ|2 ) U v)(x) = (2π)−n eix.ξ g(|ξ|2 )û(ξ)dξ ∀v ∈ L2 (Rn ).

The following operators are of special interest:


- The heat semigroup e−t∆ , t ≥ 0.
- The wave group eit∆ , t ∈ R.
- The complex powers ∆z , <z ≥ 0.
These operators can be similarly defined by Borel functional calculus for any self-
adjoint unbounded operators with nonnegative spectrum.

3.4. Compact Operators


In the sequel for any r > 0 we denote by B(0, r) the (open) unit ball of H of
radius r about the origin.
Definition 3.4.1. An operator T ∈ L(H) is said to be compact when T (B(0, 1)
is compact in H.
We denote by K the space of compact operators.
The following lemma will be useful to study compact operators.
Proposition 3.4.2. Let T ∈ L(H). Then the following are equivalent:
(i) T is a compact operator.
(ii) For any bounded sequence (ξn )n≥0 ⊂ H there is a subsequence (ξnk )k≥0
such that the sequence (T ξnk )k≥0 converges in norm.
(iii) For any sequence (ξn )n≥0 ⊂ H converging weakly to 0 the sequence (T ξn )n≥0
converges to 0 in norm.
7
(iv) There is an orthonormal basis (ξn )n≥0 of H such that
lim kT|EN⊥ k = 0,
N →∞

where we have denoted by EN the span of ξ0 , . . . , ξN −1 .


(v) For any orthonormal basis (ξn )n≥0 of H,
lim kT|EN⊥ k = 0,
N →∞

where we have denoted by EN the span of ξ0 , . . . , ξN −1 .


Proof. Taking into account that the implication (v) ⇒ (iv) is immediate, to
prove the proposition we only need to prove the implications (ii) ⇒ (i), (iii) ⇒ (ii),
(iv) ⇒ (iii) and (i) ⇒ (v).
• (ii) ⇒ (i): Assume that T sastifies the condition (ii). Then for any sequence
(ξn )n≥0 ⊂ B(0, 1) the sequence (T ξn )n≥0 admits a convergent subsequence. By
virtue of the Bolzano-Weierstrass criterion this proves that T (B(0, 1)) is compact,
i.e., T is a compact operator. Thus (ii) implies (i).
• (iii) ⇒ (ii): Suppose that T satisfies the condition (iii). Let (ξn )n≥0 ⊂ H be a
bounded sequence, i.e., there exists r > 0 such that ξn ∈ B(0, r) for all n. By
Alaoglu theorem (see [Fo, pp. 169-170]) the ball B(0, r) is precompact with respect
to the weak topology and, as H is separable, the weak topology is metrizable
(see [Fo]), so by the Bolzano-Weierstrass criterion there is a subsequence (ξnk )k≥0
converging weakly to some ξ in H. Then ξnk − ξ converges weakly to 0, so by (iii)
the sequence T (ξnk − ξ) converges to 0 in norm, i.e., T ξnk → T ξ in norm. This
shows that (iii) implies (ii).
• (iv) ⇒ (iii): Suppose that T satisfies (iv). Thus there exists an orthonormal basis
(ξn )n≥0 of H such that if, for any N ∈ N, we denote by EN the span of ξ0 , . . . , ξN −1
then kT|EN⊥ → 0 as N → ∞. In addition, we denote by ΠN the orthogonal projection
onto EN .
Let (ηk )k≥0 ⊂ H be a sequence converging weakly to 0. In particular (ηk )k≥0
is weakly bounded, and hence is bounded in norm by the uniform boundedness
principle and the fact H is isometrically isomorphic to its dual (see [Fo, pp. 163,
174–175 ]). Thus there exists P C > 0 such that kηk k ≤ C for all k ∈ N0 .
Let  > 0. Since ΠN ηk = n<N hξn , ηk iξn we have

(3.8)
X
kT ηk k ≤ kT ΠN ηk k + kT (1 − ΠN )ηk k ≤ |hξn , ηk i|kT ξn k + kT (1 − ΠN )kkηk k
n<N
X
≤ kT k |hξn , ηk i| + CkT|EN⊥ k.
n<N

Since kT|EN⊥ k → 0 as N → ∞ by choosing N large enough we have


(3.9) kT (1 − ΠN )k < .
P
As ηk converges weakly to 0 we see that n<N |hξn , ηk i| goes to 0 as k → ∞, and
hence there exists k0 ∈ N0 such that, for any k ≥ k0 ,
X
(3.10) |hξn , ηk i| < .
n<N

8
Combining (3.8), (3.9) and (3.10) we see that, for all k ≥ k0 , we have
kT ηk k ≤ (ν0 + C).
This shows that T ηk → 0 in norm as k → ∞. Thus (iv) implies (iii).
• (i) ⇒ (v): Suppose that T is a compact operator. Let (ξn )n≥0 be an orthonormal
basis of H. For any N ∈ N we denote by EN the span of ξ0 , . . . , ξN −1 . Assume that
the sequence (kT|EN⊥ k)N ≥1 does not converge to 0 as N → 0. Since this is a non-
increasing sequence of non-negative numbers there is c > 0 such that kT|EN⊥ k > c

for all N ∈ N. Therefore, for every N ∈ N, there is a unit vector ηN ∈ EN such
that kT ηN k > c.

P
Let ξ ∈ H. As ηN is contained in EN , and hence ηN = n≥N hξn , ηN iξn , we
have
X X 2  X 2
2 2
|hηn , ξi| ≤ |hηN , ξn ihξ, ξn i| ≤ |hηN , ξn i| |hξ, ξn i| .
n≥N n≥N n≥N

2 2 2
P P
Since n≥N |hηN , ξn i| = kηN k = 1 and n≥N |hξ, ξn i| → 0 as N → ∞, it
follows that hηn , ξi → 0 as N → ∞. Thus ηN converges weakly to 0 as N → ∞. As
T is continuous with respect to the weak topology it follows that T ηN converges
weakly to 0 as N → ∞.
On the other hand, the sequence (T ηN )n≥1 is contained in the image by T of the
unit sphere, which is precompact since T is compact. Therefore, by the Bolzano-
Weierstrass criterion there is a subsequence (ηNk )k≥0 such that T ηNk converges in
norm to some ζ as k → ∞. As T ηNk converges weakly to 0 we must have ζ = 0,
i.e., T ηNk converges to 0 in norm. This contradicts the fact that kT ηN k > c for all
N ∈ N. Therefore, it is not possible for the sequence (kT|EN⊥ k)N ≥1 to not converge
to 0. This proves that kT|EN⊥ k → 0 as N → ∞. Thus (i) implies (v). The proof is
complete. 

Proposition 3.4.3. K is a closed two-sided ideal of L(H).


Proof. Let (ξn )n≥0 be any sequence converging weakly to 0.
For j = 1, 2 let Tj ∈ K and λj ∈ C. By Proposition 3.4.2 (iii) the sequences
(T1 ξn )n≥0 and (T2 ξn )N converge to 0 in norm, and so (λ1 T1 + λ2 T2 )ξn )n≥0 too
converges to 0 in norm. It then follows from Proposition 3.4.2 (iii) that λ1 T1 + λ2 T2
is a compact operator. Thus K is a subspace of L(H).
Let T ∈ K and let A, B ∈ L(H). Since B is continuous with respect to the
weak topology the sequence (Bξn )n≥0 converges weakly to 0. Since T is compact
Proposition 3.4.2 (iii) insures us that (T Bξn )n≥0 converges to 0 in norm. Then
(AT Bξn )n≥0 converges to 0 in norm too. Thanks to Proposition 3.4.2 (iii), this
shows that AT B is a compact operator. Therefore, we see that K is a two-sided
ideal of L(H).
It remains to show that K is closed. Thus, let (Tk )k≥0 ⊂ K be a sequence such
that Tk → T in L(H) and let us show that T is compact. Let  > 0. Then for
k large enough kT − Tk k < . Since Tk is compact by Proposition 3.4.2 (iii) the
sequence (Tk ξn )n≥0 converges to 0 in norm, and hence there exists N ∈ N, such
that kTk ξn k <  for all n ≥ N . Then, for any n ≥ N , we have
kT ξn k ≤ kTk ξn k + k(T − Tk )ξn k ≤ kTk ξn k + kT − Tk kkξn k < 2.
9
Thus (T ξn )n≥0 converges to 0 in norm, which by Proposition 3.4.2 (iii) shows that
T is compact. This proves that K is closed, completing the proof. 
Proposition 3.4.4. For any T ∈ L(H),
(3.11) T ∈ K ⇐⇒ T ∗ ∈ K ⇐⇒ |T | ∈ K.
Proof. Let T ∈ L(H) and let T = U |T | be its polar decomposition. The
fact that K is a two-sided ideal then implies that if |T | is compact, then so is T .
Likewise, as by Proposition 3.1.8 we have |T | = U ∗ T , if T is compact, then so is
|T |.
Proposition 3.1.8 also tells us that T ∗ = U T U ∗ . Therefore, we also see that
if T is compact then so is T ∗ . Upon substituting T ∗ for T , we see that if T ∗ is
compact, then T is compact too. The proof is complete. 
Proposition 3.4.5. K is a sub-C ∗ -algebra of L(H), and hence is a C ∗ -algebra.
Proof. Since Proposition 3.4.3 tells us that K is a closed two-sided ideal, we
see that K is a closed subalgebra of L(H). As by Proposition 3.4.4 K is closed
under the involution of L(H) it follows that K is a sub-C ∗ -algebra of L(H). 
Theorem 3.4.6 (Riesz-Schauder; see [RS, Thm. VI.15]). Let T ∈ K.
(1) T always contains 0 in its spectrum.
(2) If λ ∈ Sp T \ 0, then λ is an eigenvalue with finite multiplicity.
(3) Sp T is either finite or consists of a sequence of complex numbers converg-
ing to 0.
In the sequel, for any vectors ξ and η in H, we denote by ξ ⊗ η ∗ the element of
L(H) defined by
(3.12) (ξ ⊗ η ∗ )ζ := hη, ξ 0 iξ ∀ζ ∈ H.

Thus in ketbra notation ξ ⊗ η is just the operator |ξihη|. This is an operator of
rank 1. If ξ is a unit vector then ξ ⊗ ξ ∗ is the orthogonal projection onto Cξ.
Theorem 3.4.7 (Hilbert-Schmidt; see [RS, Thm. VI.16]). Let T ∈ K be nor-
mal. Then T diagonalizes in an orthonormal basis, i.e., there exists an orthonormal
basis (ξn )n≥0 of H and a sequence (λn )n≥0 ⊂ C such that
(3.13) T λn = λn ξn ∀n ∈ N0 .
This result allows us to reinterpret the Borel functional calculus for normal
compact operators as follows.
Proposition 3.4.8. Let T ∈ K be normal and let (ξn )n≥0 be an orthonormal
basis of H respect to which T is diagonal, i.e., T λn = λn ξn for all n ∈ N0 . In
addition, let f be bounded function on Sp T = {λn ; n ∈ N0 }.
(1) We have
X
(3.14) f (T ) = f (λn )(ξn ⊗ ξn∗ ),
n≥0

where the series converges strongly.


(2) f (T ) is a compact operator if and only if
lim f (λn ) = 0.
n→∞
Furthermore, in this case the series (3.14) converges in norm.
10
Proof. For any n ∈ N0 we have T ξn = λn ξn . Since (ξn )n≥0 is an orthonormal
basis wePdeduce that T ∗ ξn = λn ξn for all n ∈ N0 . Let n ∈ N0 . For any polynomial
f (z) = ajk z j z k , we have
X X
f (T ) = ajk T j (T ∗ )k ξn = ajk λjn (λn )k ξn = f (λn )ξn .
It then follows from (1.14) that, for any f ∈ C(Sp T ), we have
(3.15) f (T )ξn = λn ξn ∀n ∈ N0 .
Since Sp T = {λn ; n ∈ N0 } any function on Sp T is a Borel function. Moreover,
by Theorem 3.4.6 0 is always in Sp T and the non-zero eigenvalues of T have finite
multiplicities, i.e., each non-zero eigenvalue λ appears at most finitely many times
in the sequence (λn )n≥0 . Thus λn → 0 as n → ∞, and hence a function f (λ) on
Sp T is continuous iff limn→∞ f (λn ) = f (0). Let f be a bounded function on Sp T
and, for N ∈ N, let fN be the function on Sp T defined by

f (λn ) if λ = λn with n < N ,
fN (λ) =
f (0) if λ = 0 or λ 6= λn for all n < N .
As limn→∞ fN (λn ) = f (0) = fN (0), we see that fN is a continuous function on
Sp T , and hence it satisfies (3.15). Moreover, for all n ∈ N0 ,
|fN (λn )| ≤ kf k∞ and lim fN (λn ) = f (λn ).
N →∞

Thus (fN )N ≥1 is a bounded sequence in L∞ (Sp T ) which converges pointwise to f .


It then follows from Theorem 3.2.2 and (3.15) that, for all n ∈ N0 ,
(3.16) f (T )ξn = lim fN (T )ξn = lim fN (λn )ξn = f (λn )ξn .
N →∞ N →∞
P
Let ξ ∈ H. Since ξ = n≥0 hξn , ξiξn and f (T ) is continuous, using (3.16) we
get
X X X
f (T )ξ = hξn , ξif (ξn ) = hξn , ξif (λn )ξn = f (λn )(ξn ⊗ ξn∗ )ξ,
n≥0 n≥0 n≥0

which proves (3.14).


Next, for N ∈ N let EN the span of ξ0 , . . . , ξN1 and denote by ΠN be the
orthogonal projection onto EN . Then by Proposition 3.4.2 (iv) the operator f (T )
is compact if and only if
(3.17) kf (T )|EN⊥ k = kf (T )(1 − ΠN )k −→ 0 as N −→ 0.
Set νN = supn≥N |f (λn )|. Since f (T )ξn = f (λn )ξn for all n ∈ N0 , we see that
νN ≤ kf (T )|EN⊥ k = kf (T )(1 − ΠN )k. Conversely, let ξ ∈ H. Since (ξn )n≥0 is an
orthonormal basis, from (3.14) we get
X X
kf (T )(1 − ΠN )ξk2 = |hξn , ξi|2 |f (λn )|2 ≤ νN
2
|hξn , ξi|2 = νN
2
kξk2 .
n≥N n≥0

Thus kf (T )(1 − ΠN )k ≤ νN . Thefore, we have


(3.18) kf (T )|EN⊥ k = kf (T )(1 − ΠN )k = sup |f (λn )|.
n≥N

Combining this with the condition (3.17) then shows that f (T ) is compact if and
only if limn→∞ f (λn ) = 0.
11
Finally, using (3.14) we get
X
f (T )(1 − ΠN ) = f (λn )(ξn ⊗ ξn∗ ),
n≥N

where the series converges strongly. Thus,


X
f (λn )(ξn ⊗ ξn∗ ) = kf (T )(1 − ΠN )k = sup |f (λn )|.
n≥N
n≥N

Therefore, if limn→∞ f (λn ) = 0, then the series (3.14) converges to f (T ) in norm.


The proof is complete. 
Example 3.4.9. Let T ∈ K be normal. Then, with the notation of Proposi-
tion 3.4.8, we have
X
(3.19) T = λn (ξn ⊗ ξn∗ ),
λn 6=0

where the series converges in norm. Moreover, if we let T = U |T | be the polar


decomposition, then
X X
U= |λn |−1 λn (ξn ⊗ ξn∗ ) and |T | = |λn |(ξn ⊗ ξn∗ ),
λn 6=0 n≥0

where the first series converges strongly (unless T has finite rank, in which case it
is a finite sum), and the second series converges in norm.

Bibliography
[Fo] Folland, G.: Real analysis. 2nd edition. John Wiley & Sons, Inc., New York,
1999.
[RS] Reed, M.; Simon, B.: Methods of modern mathematical physics. I. Functional
analysis. Second edition. Academic Press, Inc., New York, 1980.

12

You might also like