0% found this document useful (0 votes)
3 views15 pages

Lecture_notes (1)-pages

This document discusses self-adjoint operators in Hilbert spaces, defining properties such as normality and unitarity, and presents propositions and theorems regarding their behavior. It introduces positive operators and their ordering, along with the existence of positive square roots for positive operators. The document concludes with the uniqueness of positive square roots and their commutation with other operators.

Uploaded by

f20220772
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)
3 views15 pages

Lecture_notes (1)-pages

This document discusses self-adjoint operators in Hilbert spaces, defining properties such as normality and unitarity, and presents propositions and theorems regarding their behavior. It introduces positive operators and their ordering, along with the existence of positive square roots for positive operators. The document concludes with the uniqueness of positive square roots and their commutation with other operators.

Uploaded by

f20220772
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/ 15

Lecture 34 : Self-adjoint operator

Definition 34.1. A bounded linear operator T on a Hilbert space H is said to be self-adjoint if


T ∗ = T , and is said to be normal if T ∗ T = T T ∗ . T is said to be unitary if T ∗ T = T T ∗ = I.

It is easy to see that if T is self-adjoint or unitary, then T is normal.

Example: Let A = [aij ]n×n be a matrix of order n with complex entries. Then A defines a linear
operator on Cn . Note that if we consider the vectors of Cn as column vectors, then for all x, y ∈ Cn ,

⟨Ax, y⟩ = y T Ax and ⟨x, A∗ y⟩ = (A∗ y)T x = y T (A∗ )T x.

This gives that


y T Ax = y T (A∗ )T x =⇒ (A∗ )T = A =⇒ A∗ = (A)T .

Proposition 34.1. Let T be a bounded linear operator on a Hilbert space H. Then the following
statements hold:

(i) If T is self-adjoint, then ⟨T x, x⟩ is real for all x ∈ H.

(ii) If H is complex and ⟨T x, x⟩ is real for all x ∈ H, then T is self-adjoint.

(iii) The product of two self-adjoint operators is self-adjoint if and only if both commute.

Proof. (i) Suppose that T is self-adjoint. Then for all x ∈ H,

⟨T x, x⟩ = ⟨x, T x⟩ = ⟨T x, x⟩.

Thus, ⟨T x, x⟩ is real.

(ii) Suppose that H is complex and ⟨T x, x⟩ is real for all x ∈ H. Then

⟨T x, x⟩ = ⟨T x, x⟩ = ⟨x, T x⟩ = ⟨T ∗ x, x⟩ for all x ∈ H.

Thus,
⟨(T − T ∗ )x, x⟩ ≥ 0 for all x ∈ H.
Since H is complex, we conclude that T = T ∗ .

(iii) Note that (ST )∗ = T ∗ S ∗ = T S. It is now obvious that ST is self-adjoint if and only if
ST = T S.

Theorem 34.1. Let (Tn )n≥1 be a sequence of bounded self-adjoint operators on a Hilbert space H.
If Tn → T as n → ∞, then T is self-adjoint.

Proof. Note that

∥T − T ∗ ∥ ≤ ∥T − Tn ∥ + ∥Tn − Tn∗ ∥ + ∥Tn∗ − T ∗ ∥ = ∥T − Tn ∥ + 0 + ∥Tn − T ∥ → 0 as n → ∞.

Thus, T ∗ = T .

Proposition 34.2. Let U be a unitary operator on a Hilbert space H. Then the following are true:

62
(i) U is an isometry, i.e., ∥U x∥ = ∥x∥ for all x ∈ H.

(ii) ∥U ∥ = 1, provided H ̸= {0}.

(iii) U −1 is unitary.

(iv) U is normal.

(v) If V is another unitary on H, then U V is unitary.

(vi) A bounded linear operator T on a complex Hilbert space is unitary if and only if T is an
isometry and surjective.

Proof. (i) ∥U x∥2 = ⟨U x, U x⟩ = ⟨U ∗ U x, x⟩ = ∥x∥2 .

(ii) Follows from (i).

(iii) Since U ∗ = U −1 , it follows that U −1 is unitary.

(iv) Already noted.

(v) Suppose that V is a unitary on H. Then

(U V )∗ U V = V ∗ U ∗ U V = V ∗ V = I.

Similarly, we can show that U V (U V )∗ = I. Thus, U V is unitary.

(vi) Suppose that T is isometry and surjective. Then

⟨T ∗ T x, x⟩ = ⟨T x, T x⟩ = ⟨x, x⟩ for all x ∈ H.

Thus, T ∗ T = I (as H is complex). Also,

T T ∗ = T T ∗ T T −1 = T T −1 = I.

Hence T is unitary. Converse is obvious.

63
Lecture 35 : Positive operator

Definition 35.1. Let T be a bounded self-adjoint operator on a Hilbert space H. We say that T is
positive if ⟨T x, x⟩ ≥ 0 for all x ∈ H. If T1 , T2 are bounded self-adjoint operators on H, then we say
that T1 ≤ T2 if T2 − T1 is positive. For a positive operator T , we write T ≥ 0.

Remark: We may define a partial ordering on the set of all bounded self-adjoint operators on a
Hilbert space. In fact, we set T1 ≤ T2 if and only if T2 − T1 ≥ 0. Further, it is easy to see that the
sum of two positive operators is positive.

Theorem 35.1. Let S and T be positive operators on a Hilbert space H. If ST = T S, then ST is


positive.

Proof. If S = 0 or T = 0, then clearly ST is positive. Assume that both S and T are non-zero.
Define a sequence (Sn )n≥1 of operators on H as follows:

S
S1 = , and Sn+1 = Sn − Sn2 for all n ≥ 1.
∥S∥

We claim that
0 ≤ Sn ≤ I for all n ≥ 1. (35.31)
We prove Eq.(35.31) by induction. For n = 1, clearly S1 ≥ 0 as S ≥ 0. Further, for all x ∈ H,

1 ∥Sx∥∥x∥ ∥S∥∥x∥2
⟨S1 x, x⟩ = ⟨Sx, x⟩ ≤ ≤ = ∥x∥2 = ⟨Ix, x⟩.
∥S∥ ∥S∥ ∥S∥

Thus, S1 ≤ I. Suppose that (1) is true for n. Note that

⟨Sn2 (I − Sn )x, x⟩ = ⟨(I − Sn )Sn x, Sn x⟩ ≥ 0 for all x ∈ H.

Thus,
Sn2 (I − Sn ) ≥ 0.
Similarly, we can also show that
Sn (I − Sn )2 ≥ 0.
Adding them, we obtain

0 ≤ Sn2 (I − Sn ) + Sn (I − Sn )2 = Sn − Sn2 = Sn+1 .

Since I − Sn ≥ 0 and Sn2 ≥ 0, we get

0 ≤ I − Sn + Sn2 = I − Sn+1 .

This proves Eq.(35.31).


Observe that

S1 = S12 + S2 = S12 + S22 + S3 = · · · = S12 + · · · + Sn2 + Sn+1 .

Since Sn+1 > 0, we get


S12 + · · · + Sn2 = S1 − Sn+1 ≤ S1 . (35.32)

64
Thus, for all x ∈ H,
n
X n
X n
X
∥Sj x∥2 = ⟨Sj x, Sj x⟩ = ⟨Sj2 x, x⟩ ≤ ⟨S1 x, x⟩.
j=1 j=1 j=1
P∞
Since n is arbitrary, we get that j=1 ∥Sj x∥2 is convergent. Hence, ∥Sn x∥ → 0 or Sn x → 0 as
n → ∞. By Eq.(35.32),
 
Xn
 Sj2  x = (S1 − Sn+1 )x → S1 x as n → ∞.
j=1

S
Note that all Sj commute with T , as they are sums and products of S1 = ∥S∥ . Thus, we get
n
X n
X
⟨ST x, x⟩ = ∥S∥⟨T S1 x, x⟩ = ∥S∥ lim ⟨T Sj2 x, x⟩ = ∥S∥ lim ⟨T Sj x, Sj x⟩ ≥ 0.
n→∞ n→∞
j=1 j=1

Thus, ST ≥ 0.

Theorem 35.2. Let (Tn )n≥1 be a sequence of bounded self-adjoint linear operators on a complex
separable Hilbert space H such that
Tn ≤ Tn+1 and Tn ≤ K for all n ≥ 1,
where K is a bounded self-adjoint operator on H. Suppose that Ti Tj = Tj Ti for all i, j ≥ 1 and
Tj K = KTj for all j ≥ 1. Then there exists a bounded self-adjoint operator T on H such that
Tn x → T x for all x ∈ H, and T ≤ K.
Proof. Let Sn = K − Tn for all n ≥ 0. Then Sn ≥ 0 for all n ≥ 1. Let m < n. Then
2
Sm − Sm Sn = (Sm − Sn )Sm = (Tn − Tm )(K − Tm ) ≥ 0
by the preceding theorem. Similarly, we get
Sn Sm − Sn2 = Sn (Sm − Sn ) = (K − Tn )(Tn − Tm ) ≥ 0.
Thus,
Sn2 ≤ Sn Sm ≤ Sm
2
for m < n.
This gives that
∥Sn x∥2 = ⟨Sn x, Sn x⟩ = ⟨Sn2 x, x⟩ ≤ ⟨Sn Sm x, x⟩ ≤ ⟨Sm
2
x, x⟩ = ∥Sm x∥2 ≥ 0.
Hence, for a fixed x, (∥Sn x∥2 )n≥1 is a monotonic decreasing sequence of non-negative numbers, and
thus convergent. Also,
∥Sn x − Sm x∥2 = ⟨(Sn − Sm )x, (Sn − Sm )x⟩ = ∥Sn x∥2 − 2⟨Sn Sm x, x⟩ + ∥Sm x∥2
≤ ∥Sm x∥2 − ∥Sn x∥2 → 0 as m, n → ∞.
Thus, (Sn x)n≥1 is Cauchy for each fixed x ∈ H. Consequently, (Tn x)n≥1 is Cauchy for each fixed
x ∈ H. Since H is complete, let y ∈ H be such that Tn x → y. Define T : H → H as
T x = lim Tn x.
n→∞

Then T is bounded by the uniform boundedness principle. Clearly, T is self-adjoint as each Sn is


self-adjoint. Since Tn ≤ K, it follows that T ≤ K.

65
Lecture 36 : Positive square root

Definition 36.1. Let T be a positive operator on a complex Hilbert space H. Then a bounded self-
adjoint operator A is said to be a square root of T if A2 = T . If A ≥ 0, then A is called a positive
square root of T and is denoted by T 1/2 .
Theorem 36.1. Let T be a positive operator on a complex Hilbert space H. Then T has a unique
positive square root A. Further, A commutes with every operator on H which commutes with T .
T
Proof. If T = 0, then A = 0. Assume that T ̸= 0. Set Q := ∥T ∥ . Then

∥T x∥
⟨Qx, x⟩ ≤ ∥x∥ ≤ ∥x∥2 = ⟨Ix, x⟩, for all x ∈ H.
∥T ∥
Thus, Q ≤ I. If B is the unique positive square root of Q, then
 2
∥T ∥1/2 B = ∥T ∥B 2 = ∥T ∥Q = T.

Thus, ∥T ∥1/2 B is the positive square root of T . Hence without loss of generality, we may assume
that T ≤ I. Define
1
A0 = 0 and An+1 = An + (T − A2n ), n ≥ 0.
2
We claim that
An ≤ An+1 and An ≤ I for all n ≥ 0. (36.33)
We prove Eq.(36.33) by induction. Clearly A0 ≤ I and A0 ≤ A1 . Assume that Eq.(36.33) is true
for n. Since I − An ≥ 0 and I − T ≥ 0, we obtain that
1 1 1
0 ≤ (I − An )2 + (I − T ) = I − An − (T − A2n ) = I − An+1 .
2 2 2
Also,
 
1 2 1 2 1
An+2 − An+1 = An+1 + (T − An+1 ) − An − (T − An ) = (An+1 − An ) I − (An + An+1 ) ≥ 0.
2 2 2
By preceding theorem, there exists a bounded self-adjoint operator A such that An x → Ax for all
x ∈ H. Note that
1
An+1 x − An x = (T x − A2n x) → 0 as n → ∞.
2
2 2
Thus, An x → T x, and hence, T x = A x for all x ∈ H. Since An ≥ A0 = 0, it follows that A ≥ 0.
Let S be an operator which commutes with T . Then SAn = An S for all n ≥ 0. Thus, SAn x = An Sx
for all x ∈ H and n ≥ 0. Letting n → ∞, we get SA = AS.
For uniqueness Let A and B be positive square roots of T . Then

A2 = B 2 = T.

Also,
BT = BB 2 = B 2 B = T B.
Hence, AB = BA. Let x ∈ H and y = (A − B)x. Then we get

⟨Ay, y⟩ + ⟨By, y⟩ = ⟨(A + B)y, y⟩ = ⟨(A2 − B 2 )x, y⟩ = 0.

66
Thus, ⟨Ay, y⟩ = 0 and ⟨By, y⟩ = 0. Let C be the positive square root of A. Then

0 = ⟨Ay, y⟩ = ⟨C 2 y, y⟩ = ⟨Cy, Cy⟩ = ∥Cy∥2 .

That is, Cy = 0 ⇒ C 2 y = Ay = 0. Similarly, By = 0, and hence (A − B)y = 0. Note that

∥Ax − Bx∥2 = ⟨(A − B)2 x, x⟩ = ⟨(A − B)y, x⟩ = 0 for all x ∈ H.

This proves that A = B.

67
Lecture 37 : Orthogonal projections

Theorem 37.1. A bounded linear operator P : H → H on a Hilbert space H is an orthogonal


projection if and only if P is self-adjoint and idempotent.
Proof. Suppose that P is an orthogonal projection onto a closed subspace Y of H. Then H = Y ⊕Y ⊥ .
If x ∈ H, then x = y + z, where y ∈ Y and z ∈ Y ⊥ . Note that
P 2 x = P (P x) = P (y) = y = P x.
Thus, P 2 = P . Further, let x1 , x2 ∈ H. Then x1 = y1 + z1 and x2 = y2 + z2 . Hence,
⟨P x1 , x2 ⟩ = ⟨y1 , y2 + z2 ⟩ = ⟨y1 , y2 ⟩ = ⟨x1 , P x2 ⟩.
This shows that P ∗ = P .
Conversely, suppose that P 2 = P and P ∗ = P . Then for each x ∈ H, we can write
x = P x + (I − P )x.
Note that for x, y ∈ H,
⟨P x, (I − P )y⟩ = ⟨x, P (I − P )y⟩ = ⟨x, 0⟩ = 0.
Thus, P (H) ⊥ (I − P )(H). Also,
(I − P )P x = P x − P 2 x = 0, for all x ∈ H.
Hence P (H) ⊆ N (I − P ). If (I − P )x = 0, then P x = x. That is, x ∈ P (H). Thus,
N (I − P ) ⊆ P (H),
Consequently, P (H) = N (I − P ). Hence, Y = P (H) is closed. Finally,
P (P x) = P 2 x = P x for all x.

Proposition 37.1. If P is an orthogonal projection on a Hilbert space H, then the following are
true:
(i) ⟨P x, x⟩ = ∥P x∥2 for all x ∈ H.
(ii) P ≥ 0.
(iii) ∥P ∥ = 1 if P (H) ̸= {0}.
Proof. For x ∈ H, note that
⟨P x, x⟩ = ⟨P x, P x⟩ = ∥P x∥2 ≥ 0.
This proves (i) and (ii). To see (iii), assume that P (H) ̸= {0}. Note that for all x ∈ H,
∥P x∥2 = ⟨P x, x⟩ ≤ ∥P x∥∥x∥ =⇒ ∥P x∥ ≤ ∥x∥ if P x ̸= 0.
Hence ∥P ∥ ≤ 1. Also ∥P x∥ = 1 if x ∈ P (H) \ {0}. This proves that
∥P ∥ = 1 if P (H) ̸= {0}.

68
Theorem 37.2. Let P1 , P2 be orthogonal projections on a Hilbert space H. Then P1 P2 is an
orthogonal projection if and only if P1 P2 = P2 P1 . In this case, P1 P2 is an orthogonal projection
onto P1 (H) ∩ P2 (H). Further, two closed subspaces M1 , M2 of H are orthogonal if and only if
the corresponding projections satisfy PM1 PM2 = 0, where PMj is an orthogonal projection onto
Mj , j = 1, 2.
Proof. Suppose that P1 P2 = P2 P1 . Let P = P1 P2 . Then clearly P ∗ = P . Also,

P 2 = (P1 P2 )(P1 P2 ) = P12 P22 = P1 P2 .

Therefore, by preceding theorem, P1 P2 is an orthogonal projection. It is left as an exercise to verify


that P1 P2 (H) = P1 (H) ∩ P2 (H).
Conversely, if P1 P2 is an orthogonal projection, then

P1 P2 = (P1 P2 )∗ = P2∗ P1∗ = P2 P1 .

Suppose that M1 ⊥ M2 . Then M1 ∩ M2 = {0}, and hence, by previous part, PM1 PM2 = 0.
Conversely, suppose that PM1 PM2 = 0. Then for every x ∈ M1 and y ∈ M2 , we get

⟨x, y⟩ = ⟨PM1 x, PM2 y⟩ = ⟨x, PM1 PM2 y⟩ = 0.

Thus, M1 ⊥ M2 .

Theorem 37.3. Let P1 , P2 be orthogonal projections on a Hilbert space H. Then the following are
true:
(i) P1 + P2 is an orthogonal projection if and only if P1 (H) ⊥ P2 (H).

(ii) If P1 + P2 is an orthogonal projection, then

(P1 + P2 )(H) = P1 (H) ⊕ P2 (H).

Proof. (i) Let P = P1 + P2 . Suppose that P is an orthogonal projection. Then we must have
P 2 = P . Hence we get

(P1 + P2 )2 = P12 + P1 P2 + P2 P1 + P22 = P1 + P2 + P1 P2 + P2 P1 .

Thus, we get
P1 P2 + P2 P1 = 0.
Multiplying by P2 from the left we get

P2 P1 P2 + P22 P1 = 0.

Multiplying by P2 from the right,

P2 P1 P2 + P2 P1 P2 = 0 ⇒ 2P2 P1 P2 = 0 ⇒ P2 P1 P2 = 0.

Putting this in above, we get P2 P1 = 0. Hence P1 (H) ⊥ P2 (H). Conversely, if P1 (H) ⊥ P2 (H),
then P1 P2 = P2 P1 = 0. Then it is clear that

(P1 + P2 )2 = P1 + P2 and (P1 + P2 )∗ = P1 + P2 .

Hence P1 + P2 is an orthogonal projection.

69
(ii) Suppose that P = P1 + P2 is an orthogonal projection. Then P x = P1 x + P2 x for all x ∈ H,
and hence, P (H) ⊆ P1 (H) ⊕ P2 (H). Let x ∈ P1 (H) ⊕ P2 (H). Then x = x1 + x2 , where
x1 ∈ P1 (H), x2 ∈ P2 (H). Thus,

P x = P1 (x1 + x2 ) + P2 (x1 + x2 ) = x1 + x2 = x.

Hence x ∈ P (H). Consequently, P (H) = P1 (H) ⊕ P2 (H).

70
Lecture 38 : Spectrum of a bounded linear operator

Definition 38.1. Let T be a bounded linear operator on a complex normed linear space X. A point
λ ∈ C is said to be a regular value of T if (T − λI)−1 exists and is bounded. The set of all regular
values of T is called the resolvent set of T and is denoted by ρ(T ). The complement of the resolvent
set of T is called the spectrum of T and is denoted by σ(T ).
The spectrum of T has several disjoint parts which are defined as follows:

Point spectrum or discrete spectrum or eigenvalues: The point spectrum of T , denoted by


σp (T ), contains all λ ∈ σ(T ) for which (T − λI)−1 does not exist. Equivalently, λ ∈ σp (T ) if and
only if (T − λI) is not one-one.

Continuous spectrum: The continuous spectrum of T , denoted by σc (T ), contains all λ ∈ σ(T )


for which the range of (T − λI) is dense, (T − λI)−1 exists but (T − λI)−1 is unbounded.

Residual spectrum: The residual spectrum of T , denoted by σr (T ), contains all λ ∈ σ(T ) for
which the range of (T − λI) is not dense but (T − λI)−1 exists which may be bounded or not.
It is easy to see that σp (T ), σc (T ) and σr (T ) are mutually disjoint subsets of σ(T ). Further,

σ(T ) = σp (T ) ∪ σc (T ) ∪ σr (T ).

Remark: The spectrum of an operator on a finite dimensional normed linear space (i.e., the spec-
trum of a matrix) is equal to the point spectrum. In case of infinite dimensional normed linear space,
the spectrum of an operator is much larger than the point spectrum, in general.

Example 1: For a bounded linear operator T , T −1 may exist but may not be bounded. For
example, consider T : ℓ2 → ℓ2 defined as
 x2 x3 xn 
T (x1 , x2 , . . .) = x1 , , , . . . , , . . . , x = (xn )n≥1 ∈ ℓ2 .
2 3 n
Then T is bounded with ∥T ∥ = 1 (verify!). Note that

T −1 (x1 , x2 , . . .) = (x1 , 2x2 , 3x3 , . . . , nxn , . . .), x = (xn )n≥1 ∈ ℓ2 .

But T −1 is not bounded (verify!).


Lemma 38.1. Suppose that T is a bounded linear operator on a complex Banach space X. If
∥T ∥ < 1, then the operator (I − T )−1 exists and is a bounded linear operator on X. Further,

X
−1
(I − T ) = Tj = I + T + T2 + ···
j=0

Proof. Note that ∥T ∥ < 1, hence



X ∞
X
j
∥T ∥ ≤ ∥T ∥j < ∞.
j=0 j=0

Now for all m > 0,

(I − T ) I + T + T 2 + · · · + T m = I − T m+1 → 0 as m → ∞.


71
Thus, we get
(I − T ) I + T + T 2 + · · · = I + T + T 2 + · · · (I − T ) = I
 

Thus, (I − T )−1 is bounded and is given by (I − T )−1 = ∞ j


P
j=0 T .

Example 2: Consider T : ℓ2 → ℓ2 defined as

T (x1 , x2 , . . .) = (0, x1 , x2 , . . .), x = (xn )n≥1 ∈ ℓ2 .

Before proceeding towards the computation of σ(T ), let us observe few things. For any bounded
linear operator S on a Hilbert space H,

ker S ∗ = (Range(S)) . (1)

In fact, let y ∈ ker S ∗ . Then for all x ∈ H,

⟨Sx, y⟩ = ⟨x, S ∗ y⟩ = 0 ⇒ Range(S) ⊥ ker S ∗ ,



This gives ker S ∗ ⊆ Range(S) . Conversely, suppose that y ⊥ Range(S). In particular, y ⊥
Range(S). Thus, for all x ∈ H,
0 = ⟨y, Sx⟩ = ⟨S ∗ y, x⟩.
Hence, S ∗ y = 0, which proves (1).
The Hilbert space adjoint T ∗ of T is given by

T ∗ (x1 , x2 , . . .) = (x2 , x3 , x4 , . . .).

To see this, note that for all x, y ∈ ℓ2 , we get

⟨T ∗ x, y⟩ = ⟨x, T y⟩ = ⟨(x1 , x2 , . . .), (0, y1 , y2 , . . .)⟩ = x2 y1 + x3 y2 + · · · = ⟨(x2 , x3 , . . .), y⟩.

Thus, T ∗ x = (x2 , x3 , x4 , . . .).


Note that ∥T ∥ = 1, and hence for |λ| > 1,
 
T
(T − λI) = λ −I
λ

is invertible and (T − λI)−1 is bounded as T


λ < 1. Thus,

σ(T ) ⊆ {λ ∈ C : |λ| ≤ 1}.

Suppose λ ∈ σp (T ). Clearly, λ ̸= 0 as T is injective (in fact, it is isometry). Further, for any


0 ̸= λ ∈ C, if (T − λI)x = 0, then

(0, x1 , x2 , . . .) = (λx1 , λx2 , . . .) =⇒ x1 = x2 = · · · = 0.

Thus, (T − λI) is one-one for all λ ∈ C. Hence σp (T ) = ∅.


Let λ ∈ σp (T ∗ ). Then

T ∗ x = λx for some x ∈ ℓ2 =⇒ (x2 , x3 , x4 , . . .) = (λx1 , λx2 , λx3 , . . .)

We get
2 3 n
x2 = λx1 , x3 = λ x1 , x4 = λ x1 , · · · , xn+1 = λ x1 , · · ·

72
Clearly,

X
|xn |2 < ∞ if and only if |λ| < 1.
n=1

Thus, for |λ| < 1,


ker(T ∗ − λI) ̸= {0},
which gives that

Range(T − λI) ̸= {0}.
Hence (T − λI)−1 exists but Range(T − λI) is not dense. Thus,

σr (T ) = {λ ∈ C : |λ| < 1} .

We will see later that σ(T ) is a compact subset of C. Hence,

σc (T ) = {λ ∈ C : |λ| = 1}

and thus,
σ(T ) = {λ ∈ C : |λ| ≤ 1} .

73
Lecture 39 : Spectrum of a bounded linear operator continued

Theorem 39.1. Let T be a bounded linear operator on a complex Banach space X. Then ρ(T ) is
an open subset of C. Consequently, σ(T ) is a closed subset of C.

Proof. If ρ(T ) = ∅, it is open. Assume that ρ(T ) ̸= ∅. For a fixed λ0 ∈ ρ(T ) and any λ ∈ C, we have

T − λI = T − λ0 I − (λ − λ0 )I = (T − λ0 I) I − (λ − λ0 )(T − λ0 I)−1
 

Note that I − (λ − λ0 )(T − λ0 I)−1 is invertible if ∥(λ − λ0 )(T − λ0 I)−1 ∥ < 1. That is,
1
|λ − λ0 | < .
∥(T − λ0 I)−1 ∥

Thus, if this holds, T − λI is invertible. Consequently, ρ(T ) is open.

Corollary: The map λ 7→ (T − λI)−1 is analytic on ρ(T ).

Theorem 39.2. The spectrum σ(T ) of a bounded linear operator T on a complex Banach space X
is a non-empty compact set and lies in the disc {λ ∈ C : |λ| ≤ ∥T ∥}.
∥T ∥
Proof. We first show that σ(T ) is compact. To see this, let |λ| > ∥T ∥. Then |λ| < 1. Hence (I − Tλ )
is invertible, and thus, (T − λI) is invertible. Therefore,

{λ ∈ C : |λ| > ∥T ∥} ⊆ ρ(T ) =⇒ σ(T ) ⊆ {λ ∈ C : |λ| ≤ ∥T ∥} .

Thus, σ(T ) is bounded. Since σ(T ) is closed by the preceding theorem, it follows that σ(T ) is
compact.
Now, on contrary, suppose that σ(T ) = ∅. Then ρ(T ) = C. Hence the map λ 7→ (T − λI)−1 is
an entire function. Further, for |λ| > 2∥T ∥, we get
∞  j
−1 1 X ∥T ∥ 1 1
∥(T − λI) ∥≤ = ≤ .
|λ| |λ| |λ| − ∥T ∥ ∥T ∥
j=0

Thus, the map λ 7→ (T − λI)−1 is a bounded entire function, and hence by Liouville’s theorem is
constant. This is a contradiction. Hence σ(T ) is non-empty.

Definition 39.1. Let T be a bounded linear operator on a complex Banach space X. Then sup{|λ| :
λ ∈ σ(T )} is called as the spectral radius of T and is denoted by r(T ).

Gelfand’s Theorem. Let T be a bounded linear operator on a complex Banach space X. Then

r(T ) = lim ∥T n ∥1/n .


n→∞

74
Lecture 40 : Compact operators

Definition 40.1. Let X and Y be normed linear spaces and T : X → Y be a linear operator. Then
T is said to be a compact linear operator if for every bounded subset M of X, T (M ) is compact.

Example: Let X be an infinite dimensional normed linear space. Then the identity operator I on
X is not compact. In fact, take M = {x ∈ X : ∥x∥ ≤ 1}. Then we have already seen that M is not
compact. Hence I(M ) = M = M is not compact.

Proposition 40.1. Every compact linear operator is bounded

Proof. Let T : X → Y be a compact linear operator and M = {x ∈ X : ∥x∥ ≤ 1}. Then T (M ) is


compact, and hence, bounded. Thus,

∥T ∥ = sup ∥T x∥ < ∞.
∥x∥=1

Thus, T is bounded.

Theorem 40.1 (Compactness criterion). Let X and Y be normed linear spaces and T : X → Y
be a linear operator. Then T is compact if and only if it maps every bounded sequence in X onto a
sequence in Y which has a convergent subsequence.

Proof. Suppose that T is compact and (xn )n≥1 is a bounded sequence in X. Then {T xn : n ≥ 1} is
compact in Y , and therefore, sequentially compact. Hence (T xn )n≥1 has a convergent subsequence.
Conversely, every bounded sequence (xn )n≥1 in X has a subsequence (xnk )k≥1 such that (T xnk )k≥1
is convergent. Let M be any bounded subset of X and (T xn )n≥1 be any sequence in T (M ). Clearly,
(xn )n≥1 ⊂ M is bounded as M is bounded. Therefore, by assumption, there exists a subsequence
(T xnk ) which is convergent (in T (M )). Since T (M ) is dense in T (M ), it follows that T (M ) is
sequentially compact, and hence, T (M ) is compact.

Theorem 40.2. Let X and Y be normed linear spaces and T : X → Y be a linear operator. Then
the following statements are true:

(i) If T is bounded and dim T (X) < ∞, then T is compact.

(ii) If dim X < ∞, then T is compact.

Proof. (i) Suppose that T is bounded and dim T (X) < ∞. Let (xn )n≥1 be any bounded sequence
in X. Then
∥T xn ∥ ≤ ∥T ∥∥xn ∥ for all n ≥ 1.
Hence (T xn )n≥1 is a bounded sequence in T (X), which is a finite-dimensional subspace. There-
fore, {T xn : n ≥ 1} is compact, and by sequential compactness, (T xn )n≥1 has a convergent
subsequence. Thus, T is compact.

(ii) If dim X < ∞, then dim T (X) < ∞. Thus, the result follows from (i).

Theorem 40.3. Let (Tn )n≥1 be a sequence of compact operators from a normed linear space X into
a Banach space Y . If ∥Tn − T ∥ → 0 as n → ∞ for some linear operator T : X → Y , then T is
compact.

75
Proof. Let (xm )m≥1 be any bounded sequence in X. Since T1 is compact, (T1 x1,m )m≥1 is convergent,
where (x1,m )m≥1 is a subsequence of (xm )m≥1 . Again, (x1,m )m≥1 is bounded and T2 is compact, so
(T2 x2,m )m≥1 is convergent, where (x2,m )m≥1 is a subsequence of (x1,m )m≥1 . Continuing in this way,
we get that the sequence (ym )m≥1 = (xm,m )m≥1 is a subsequence of (xm )m≥1 such that for every
n ≥ 1, (Tn ym )m≥1 is convergent, and hence, Cauchy in Y .
Since (xm )m≥1 is bounded, (ym )m≥1 is bounded too. Let ∥ym ∥ ≤ c for all m ≥ 1. Since
∥Tn − T ∥ → 0, there exists n0 ∈ N such that ∥Tn0 − T ∥ < ε. Since (Tn0 ym )m≥1 is Cauchy, there
exists n1 ∈ N such that
∥Tn0 yj − Tn0 yk ∥ < ε for all j, k > n1 .
Hence, for j, k ≥ n1 , we get

∥T yj − T yk ∥ ≤ ∥T yj − Tn0 yj ∥ + ∥Tn0 yj − Tn0 yk ∥ + ∥Tn0 yk − T yk ∥


≤ ∥T − Tn0 ∥∥yj ∥ + ε + ∥Tn0 − T ∥∥yk ∥
≤ εc + ε + εc = (2c + 1)ε.

Thus, (T yj )j≥1 is a Cauchy sequence in Y . Since Y is complete, (T yj )j≥1 is convergent. That is,
(T xm )m≥1 has a convergent subsequence. Hence, T is compact.

Example: Let T : ℓ2 → ℓ2 be defined as


 x2 x3 xn 
T (x1 , x2 , . . .) = x1 , , , . . . , , . . . , x = (xn )n≥1 ∈ ℓ2 .
2 3 n
Then T is compact. To see this, define
 x2 x3 xn 
Tn (x1 , x2 , . . .) = x1 , , , . . . , , 0, 0, . . . , x = (xn )n≥1 ∈ ℓ2 , n ≥ 1.
2 3 n
Then each Tn is a bounded linear compact operator. Further,

X |xj |2 1
∥(T − Tn )x∥2 = ≤ ∥x∥2 .
(j + 1)2 (n + 1)2
j=n+1

Thus, we get
1
∥T − Tn ∥ ≤ → 0 as n → ∞.
n+1
By preceding theorem, T is compact.

76

You might also like