OPERATOR THEORY ON HILBERT SPACE. Class notes. John Petrovic

Similar documents
CHAPTER VIII HILBERT SPACES

Spectral Theory, with an Introduction to Operator Means. William L. Green

Finite-dimensional spaces. C n is the space of n-tuples x = (x 1,..., x n ) of complex numbers. It is a Hilbert space with the inner product

I teach myself... Hilbert spaces

1 Functional Analysis

Chapter 8 Integral Operators

Functional Analysis. Martin Brokate. 1 Normed Spaces 2. 2 Hilbert Spaces The Principle of Uniform Boundedness 32

Elementary linear algebra

Real Analysis Notes. Thomas Goller

Functional Analysis. Franck Sueur Metric spaces Definitions Completeness Compactness Separability...

SPECTRAL THEOREM FOR COMPACT SELF-ADJOINT OPERATORS

REAL AND COMPLEX ANALYSIS

3 (Due ). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

MAT 578 FUNCTIONAL ANALYSIS EXERCISES

08a. Operators on Hilbert spaces. 1. Boundedness, continuity, operator norms

CHAPTER X THE SPECTRAL THEOREM OF GELFAND

Analysis Preliminary Exam Workshop: Hilbert Spaces

l(y j ) = 0 for all y j (1)

Spectral theory for compact operators on Banach spaces

Hilbert Spaces. Hilbert space is a vector space with some extra structure. We start with formal (axiomatic) definition of a vector space.

MTH 503: Functional Analysis

Vector Spaces. Vector space, ν, over the field of complex numbers, C, is a set of elements a, b,..., satisfying the following axioms.

Kernel Method: Data Analysis with Positive Definite Kernels

4 Hilbert spaces. The proof of the Hilbert basis theorem is not mathematics, it is theology. Camille Jordan

PROBLEMS. (b) (Polarization Identity) Show that in any inner product space

Math Linear Algebra II. 1. Inner Products and Norms

Real Analysis, 2nd Edition, G.B.Folland Elements of Functional Analysis

CHAPTER V DUAL SPACES

Infinite-dimensional Vector Spaces and Sequences

Functional Analysis II held by Prof. Dr. Moritz Weber in summer 18

HILBERT SPACES AND THE RADON-NIKODYM THEOREM. where the bar in the first equation denotes complex conjugation. In either case, for any x V define

LECTURE 7. k=1 (, v k)u k. Moreover r

Hilbert spaces. 1. Cauchy-Schwarz-Bunyakowsky inequality

MAT 570 REAL ANALYSIS LECTURE NOTES. Contents. 1. Sets Functions Countability Axiom of choice Equivalence relations 9

David Hilbert was old and partly deaf in the nineteen thirties. Yet being a diligent

Elements of Positive Definite Kernel and Reproducing Kernel Hilbert Space

NOTES ON FRAMES. Damir Bakić University of Zagreb. June 6, 2017

C.6 Adjoints for Operators on Hilbert Spaces

October 25, 2013 INNER PRODUCT SPACES

MAT 449 : Problem Set 7

Overview of normed linear spaces

Compact operators on Banach spaces

Problem Set 2: Solutions Math 201A: Fall 2016

Notes on Banach Algebras and Functional Calculus

Functional Analysis I

Problem Set 6: Solutions Math 201A: Fall a n x n,

Real Variables # 10 : Hilbert Spaces II

1 Math 241A-B Homework Problem List for F2015 and W2016

A Brief Introduction to Functional Analysis

Chapter 2 Metric Spaces

THE PROBLEMS FOR THE SECOND TEST FOR BRIEF SOLUTIONS

2) Let X be a compact space. Prove that the space C(X) of continuous real-valued functions is a complete metric space.

Topological properties

2 (Bonus). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

Recall that any inner product space V has an associated norm defined by

Commutative Banach algebras 79

1 Definition and Basic Properties of Compa Operator

WEIGHTED SHIFTS OF FINITE MULTIPLICITY. Dan Sievewright, Ph.D. Western Michigan University, 2013

Banach Spaces II: Elementary Banach Space Theory

E.7 Alaoglu s Theorem

Exercise Solutions to Functional Analysis

MATH 113 SPRING 2015

About Grupo the Mathematical de Investigación Foundation of Quantum Mechanics

1. Subspaces A subset M of Hilbert space H is a subspace of it is closed under the operation of forming linear combinations;i.e.,

CHAPTER II HILBERT SPACES

Honours Analysis III

Your first day at work MATH 806 (Fall 2015)

Functional Analysis Exercise Class

COMPACT OPERATORS. 1. Definitions

Topology. Xiaolong Han. Department of Mathematics, California State University, Northridge, CA 91330, USA address:

SPECTRAL THEOREM FOR SYMMETRIC OPERATORS WITH COMPACT RESOLVENT

Topological vectorspaces

FUNCTIONAL ANALYSIS: NOTES AND PROBLEMS

Metric Spaces and Topology

Math Solutions to homework 5

Functional Analysis F3/F4/NVP (2005) Homework assignment 2

Measurable functions are approximately nice, even if look terrible.

Chapter 7: Bounded Operators in Hilbert Spaces

THEOREMS, ETC., FOR MATH 515

Chapter 2 Linear Transformations

Numerical Linear Algebra

Real Analysis Math 131AH Rudin, Chapter #1. Dominique Abdi

TOEPLITZ OPERATORS. Toeplitz studied infinite matrices with NW-SE diagonals constant. f e C :

Math 209B Homework 2

Stone-Čech compactification of Tychonoff spaces

Part III. 10 Topological Space Basics. Topological Spaces

16 1 Basic Facts from Functional Analysis and Banach Lattices

1/12/05: sec 3.1 and my article: How good is the Lebesgue measure?, Math. Intelligencer 11(2) (1989),

Short note on compact operators - Monday 24 th March, Sylvester Eriksson-Bique

Math 350 Fall 2011 Notes about inner product spaces. In this notes we state and prove some important properties of inner product spaces.

2 Garrett: `A Good Spectral Theorem' 1. von Neumann algebras, density theorem The commutant of a subring S of a ring R is S 0 = fr 2 R : rs = sr; 8s 2

RIESZ BASES AND UNCONDITIONAL BASES

An introduction to some aspects of functional analysis

Lecture Notes in Functional Analysis

Hilbert space methods for quantum mechanics. S. Richard

CHAPTER 3. Hilbert spaces

Real Analysis Problems

FUNCTIONAL ANALYSIS LECTURE NOTES: ADJOINTS IN HILBERT SPACES

CONTENTS. 4 Hausdorff Measure Introduction The Cantor Set Rectifiable Curves Cantor Set-Like Objects...

Lecture Notes in Advanced Calculus 1 (80315) Raz Kupferman Institute of Mathematics The Hebrew University

Transcription:

OPERATOR THEORY ON HILBERT SPACE Class notes John Petrovic

Contents Chapter 1. Hilbert space 1 1.1. Definition and Properties 1 1.2. Orthogonality 3 1.3. Subspaces 7 1.4. Weak topology 9 Chapter 2. Operators on Hilbert Space 13 2.1. Definition and Examples 13 2.2. Adjoint 15 2.3. Operator topologies 17 2.4. Invariant and Reducing Subspaces 20 2.5. Finite rank operators 22 2.6. Compact Operators 23 2.7. Normal operators 27 Chapter 3. Spectrum 31 3.1. Invertibility 31 3.2. Spectrum 34 3.3. Parts of the spectrum 38 3.4. Spectrum of a compact operator 40 3.5. Spectrum of a normal operator 43 iii

iv CONTENTS Chapter 4. Invariant subspaces 47 4.1. Compact operators 47 4.2. Line integrals 49 4.3. Invariant subspaces for compact operators 52 4.4. Normal operators 56 Chapter 5. Spectral radius algebras 64 5.1. Compact operators 64

CHAPTER 1 Hilbert space 1.1. Definition and Properties In order to define Hilbert space H we need to specify several of its features. First, it is a complex vector space the field of scalars is C (complex numbers). [See Royden, p. 217.] Second, it is an inner product space. This means that there is a complex valued function x, y defined on H H with the properties that, for all x, y, z H and α, β C: (a) αx + βy, z = α x, z + β y, z ; it is linear in the first argument; (b) x, y = y, x ; it is Hermitian symmetric; (c) x, x 0; it is non-negative; (d) x, x = 0 iff x = 0; it is positive. In every inner product space it is possible to define a norm as x = x, x 1/2. Exercise 1.1.1. Prove that this is indeed a norm. Finally, Hilbert space is complete in this norm (meaning: in the topology induced by this norm). Example 1.1.1. C n is an inner product space with x, y = n k=1 x ky k and, consequently, the norm x = n k=1 x k 2. Completeness: if {x (k) } k=1 is a Cauchy sequence in Cn (here x (k) = (x (k) 1, x(k) 2,..., x(k) n )) then so is x (k) m for any fixed m, 1 m n, and C is complete. Example 1.1.2. Let H 0 denote the collection of all complex sequences, i.e. functions a : N C, characterized by the fact that a n 0 for a finite number of positive integers n. Define the inner product on H 0 by a, b = n=0 a nb n. The space H 0 is not complete in the induced norm. Indeed, the sequence {a (k) } k N, defined by a (k) n = 1/2 n if n k and a (k) n = 0 if n > k is a Cauchy sequence, but not convergent. 1

2 1. HILBERT SPACE Example 1.1.3. Let l 2 denote the collection of all complex sequences a = {a n } n=1 such that n=1 a n 2 converges. Define the inner product on l 2 by a, b = n=1 a nb n. Suppose that {a (k) } k=1 is a Cauchy sequence in l 2. Then so is {a (k) n } k=1 for each n, hence there exists a n = lim k a (k) n. First we show that a l 2. Indeed, choose K so that for k K we have a (k) a (K) 1. Then, using Minkowski s Inequality for sequences (see Royden, p. 122), for any N N, { N } 1/2 { N } 1/2 a n 2 a n a (K) n 2 + n=1 n=1 { N 1=0 a (K) n 2 } 1/2 = lim k { N n=1 a (k) n lim sup a (k) a (K) + a (K) 1 + a (K). k a (K) n 2 } 1/2 + { N n=1 } 1/2 a (K) n 2 Thus a = {a n } l 2. Moreover, {a (k) } converges to a, i.e. lim k a a (k) = 0. Let ɛ > 0 and choose M so that k, j M implies that a (k) a (j) < ɛ. For such k M and any N, we have N n=1 a n a (k) n 2 = lim N j n=1 a (j) n a (k) n 2 lim sup a (j) a (k) 2 ɛ 2. j Since N is arbitrary, it follows that a a (k) ɛ and, therefore, l 2 is Hilbert space. Example 1.1.4. The space L 2 of functions f : X C, such that X f 2 dµ < (where X is usually [0, 1] and µ Lebesgue measure). The inner product is defined by f, g = X fg dµ and L2 is complete by the Riesz Fisher Theorem (see Royden, p. 125). Example 1.1.5. The space H 2. Let X = T (the unit circle) and µ the normalized Lebesgue measure on T. The Hardy space H 2 consists of those functions in L 2 (T) such that f, e int = 0 for n = 1, 2,.... Some important facts. Proposition 1.1.1 (Parallelogram Law). x + y 2 + x y 2 = 2 x 2 + 2 y 2. Proposition 1.1.2 (Polarization Identity). 4 x, y = x + y, x + y x y, x y + i x + iy, x + iy i x iy, x iy. Exercise 1.1.2. Prove Propositions 1.1.1 and 1.1.2.

1.2. ORTHOGONALITY 3 Problem 1. Let be a norm on Banach space X, and define x, y as in Polarization Identity. Assuming that the norm satisfies the Parallelogram Law, prove that x, y defines an inner product. 1.2. Orthogonality In Linear Algebra a basis of a vector space is defined as a minimal spanning set. In Hilbert space such a definition is not very practical. It is hard to speak of minimality when a basis can be infinite. In fact, a basis can be uncountable, so if {e i } i I is such a basis, what is the meaning of i I x ie i? Definition 1.2.1. An orthonormal subset of Hilbert space H is a set E such that (a) e = 1, for all e E; (b) if e 1, e 2 E and e 1 e 2 then e 1, e 2 = 0. An orthonormal basis in H is a maximal orthonormal set. We use abbreviations o.n.s. and o.n.b. for orthonormal set and orthonormal basis, respectively. Theorem 1.2.1. Every Hilbert space has an orthonormal basis. Proof. Let e be a unit vector in H. Then E = {e} is an orthonormal set. Let M be the collection of all orthonormal sets in H that contain E. By the Hausdorff Maximal Principle (Royden, p.25) there exists a maximal chain C of such orthonormal sets, partially ordered by inclusion. Let N be the union of all elements of C. Then N is a maximal orthonormal set, hence a basis of H. If the set {e} is replaced by any orthonormal set, the same proof yields a stronger result. Theorem 1.2.2. Every orthonormal set in Hilbert space can be extended to an orthonormal basis. Example 1.2.1. For k N, let e k denote the sequence with only one non-zero entry, lying in the kth position and equal to 1. The set {e k } k N is an o.n.b. for l 2. (If a vector x l 2 is orthogonal to all e k, then each of its components is zero, so x = 0.) Example 1.2.2. The set {e 1, e 3, e 5,... } is an orthonormal set in l 2 but not a basis. Example 1.2.3. The set { 1 2π, cos t π, sin t π, cos 2t π, } sin 2t,... is an o.n.b. in L 2 ( π, π). π

4 1. HILBERT SPACE Example 1.2.4. The set { } 1 e int : n Z is another o.n.b. in L 2 ( π, π). 2π In Linear Algebra, if {e i } i I is an o.n.b. then every vector x can be written as i I x, e i e i. In Hilbert space our first task is to make sense of this sum since the index set I need not be countable. x 2. Theorem 1.2.3 (Bessel s Inequality). Let {e i } k i=1 be an o.n.s. in H, and let x H. Then k i=1 x, e i 2 Proof. If we write x i = x, e i, then k k k k k k 0 x x i e i 2 = x x i e i, x x i e i = x 2 2Re x, x i e i + x i e i, x j e j i=1 i=1 i=1 i=1 i=1 j=1 k k k k k k = x 2 2Re x i x, e i + x i x j e i, e j = x 2 2Re x i x i + x i x i = x 2 x i 2. i=1 i=1 j=1 i=1 i=1 i=1 Corollary 1.2.4. Let E = {e i } i I be an o.n.s. in H, and let x H. Then x, e i 0 for at most a countable number of i I. Proof. Let x H be fixed and let E n = {e i : x i 1/n}. If e i1, e i2,..., e ik E n then k x 2 x ij 2 k(1/n 2 ). j=1 So, for each n N, E n is a finite set, and E = n E n. In view of Corollary 1.2.4 the expressions like x, e i e i turn out to be the usual infinite series. Our next task is to establish their convergence. The following Lemma will be helpful in this direction. Lemma 1.2.5. If {x i } i N is a sequence of complex numbers and {e i } i N is an o.n.s. in H, then the series i N x ie i and i N x i 2 are equiconvergent.

1.2. ORTHOGONALITY 5 Proof. Let s m and σ m denote the partials sums of i N x ie i and i N x i 2, respectively. Then s m s n 2 = m i=n+1 x ie i 2 = m i=n+1 x ie i, m j=n+1 x je j = m i=n+1 x i 2 = σ m σ n so the series are equiconvergent. Now we can establish the convergence of i I x, e i e i. We will use notation x i = x, e i for the Fourier coefficients of x H relative to the fixed basis {e i } i I. Corollary 1.2.6 (Parseval s Identity). Let {e i } i I be an o.n.s. in H, and let x H. Then the series i I x ie i and i I x i 2 converge and i I x ie i 2 = i I x i 2. Proof. Since only a countable number of terms in each series is non-zero, we can rearrange them and consider the series i=1 x ie i and i=1 x i 2. The latter series converges by the Bessel s Inequality and Lemma 1.2.5 implies that the former series converges too. Moreover, their partial sums s m and σ m satisfy s m = σ m, so the last assertion of the corollary follows by letting m go to. Now we are in the position to show that, in Hilbert space, every o.n.b. indeed spans H. Of course, the minimality is a direct consequence of the definition. Theorem 1.2.7. Let E = {e i } i I be an o.n.b. in H. Then, for each x H, x = i I x ie i, where x i = x, e i. Proof. Let x i = x, e i and y = x i I x ie i. (Well defined since the series converges.) Then y, e k = x, e k i I x ie i, e k = 0, for each k I, so y E. If y 0, then E {y/ y } is an o.n.s., contradiciting the maximality of E, so y = 0. The following is the analogue of a well known Linear Algebra fact. We use notation card I for the cardinal number of the set I. Theorem 1.2.8. Any two orthonormal bases {e i } i I and {f j } j J in H have the same cardinal number. Proof. We will assume that both cardinal numbers are infinite. If either of them is finite, one knows from Linear Algebra that the other one is finite and equal to the first. Let j J be fixed and let I j = {i I : f j, e i

6 1. HILBERT SPACE 0}. By Corollary 1.2.4, I j is at most countable. Further, j J I j = I. Indeed, if i 0 I \ j J I j then f j, e i0 = 0 for all j J so it would follow that e i0 = 0. Since card I j ℵ 0 we see that card I card J ℵ 0 = card J. Similarly, card J card I. By Cantor Bernstein Theorem, (see, e.g., Proofs from the book, p.90) card I = card J. of H. Definition 1.2.2. The dimension of Hilbert space H, denoted by dim H, is the cardinal number of a basis In this course we will assume that dim H ℵ 0. Exercise 1.2.1. If H is an infinite dimensional Hilbert space, then H is separable iff dim H = ℵ 0. [Given a countable basis, use rational coefficients. Given a countable dense set, approximate each element of a basis close enough to exclude all other basis elements.] Next, we want to address the question: when can we identify two Hilbert spaces? We need a vector space isomorphism (i.e., a linear bijection) that preserves the inner product. Definition 1.2.3. If H and K are Hilbert spaces, an isomorphism is a linear surjection U : H K such that, for all x, y H, Ux, Uy = x, y. In this situation we say that H and K are isomorphic. Exercise 1.2.2. Prove that Ux, Uy = x, y for all x, y H iff Ux = x for all x H. Conclude that a Hilbert space isomorphism is injective. Theorem 1.2.9. Every separable Hilbert space of infinite dimension is isomorphic to l 2. Every Hilbert space of finite dimension n is isomorphic to C n. Proof. We will assume that H is an infinite dimensional Hilbert space and leave the finite dimensional case as an exercise. Since H is separable, there exists an o.n.b. {e n } n=1. For x H, let x i = x, e i and U(x) = (x 1, x 2, x 3,... ). By Parseval s Identity, the series i=1 x i 2 converges, so the sequence (x 1, x 2, x 3,... ) belongs to l 2. Thus U is well-defined, linear (because the inner product is linear in the first argument), and isometric:

1.3. SUBSPACES 7 Ux 2 = i=1 x i 2 = x 2. Finally, if (y 1, y 2, y 3,... ) l 2 then i=1 y i 2 converges so, by Lemma 1.2.5, n=1 y ne n converges and U( n=1 y ne n ) = (y 1, y 1, y 1,... ). Thus, U is surjective and the theorem is proved. Exercise 1.2.3. Prove that every Hilbert space of finite dimension n is isomorphic to C n. Problem 2. Let H ne a separable Hilbert space and M a subspace of H. Prove that M is a separable Hilbert space. Problem 3. The Haar system {ϕ m,n }, m N, 1 n 2 m, is defined as: Prove that this system is an o.n.b. of L 2 [0, 1]. 2 m/2, if n 1 2 m x n 1/2 2 m, ϕ m,n (x) = 2 m/2, if n 1/2 2 m x n 2 m, [ ) n 1 0, if x / 2 m, n 2 m. 1.3. Subspaces Example 1.3.1. Let H = L 2 [0, 1] and let G be a measurable subset of [0, 1]. Denote by L 2 (G) the set of functions in L 2 that vanish outside of G. Then L 2 (G) is a closed subspace of H. Further, if f L 2 (G) and g L 2 (G c ), then f, g = 0. Definition 1.3.1. If M is a closed subspace of the Hilbert space H, then the orthogonal complement of M, denoted M, is the set of vectors in H orthogonal to every vector in M. Exercise 1.3.1. Prove that M is a closed subspace of H. Theorem 1.3.1. Let M be a closed subspace of Hilbert space H, and let x H. Then there exist unique vectors y in M and z in M so that x = y + z.

8 1. HILBERT SPACE Proof. Let {e i } i I and {f j } j J be orthonormal bases for M and M, respectively. Their union is an o.n.b. of H so x = i I x, e i e i + j J x, f j f j and we define y = i I x, e i e i, z = j J x, f j f j. Then y M, z M, and x = y + z. Suppose now that x = y 1 + z 1 = y 2 + z 2, where y 1, y 2 M and z 1, z 2 M. Then y 1 y 2 = z 2 z 1 belongs to both M and M, so y 1 y 2, y 1 y 2 = 0 and it follows that y 1 = y 2, and consequently z 1 = z 2. Definition 1.3.2. In the situation described in Theorem 1.3.1 we say that H is the orthogonal direct sum of M and M, and we write H = M M. When z = x + y with x M and y M we often write z = x y. Theorem 1.3.1 allows us to define a map P : H M by P x = y. It is called the orthogonal projection of H onto M, and it is denoted by P M. Here are some of its properties. Theorem 1.3.2. Let M be a closed subspace of Hilbert space H and let P be the orthogonal projection on M. Then: (a) P is a linear transformation; (b) P x x, for all x H; (c) P 2 = P ; (d) Ker P = M and Ran P = M. Proof. Let {e i } i I and {f j } j J be orthonormal bases for M and M, respectively, and let Q = I P be the orthonormal projection on M. If x, x H and α, α C, then P (α x +α x ) = i I α x +α x, e i = α P x + α P x, so (a) holds. (b) If x H, then x = P x + Qx and P x Qx. Therefore, x 2 = P x 2 + Qx 2 P x 2. (c) If y M then P y = y. Now, for any x H, P x M so P 2 x = P (P x) = P x. (d) If P x = 0 then x = Qx M. If x M then Qx = x by (c), so P x = 0. The other assertion is obvious. Problem 4. Prove that P M x is the unique point in M that is nearest to x, meaning that x P M x = inf{ x h : h M}.

1.4. WEAK TOPOLOGY 9 Problem 5. In L 2 [0, 1] find the orthogonal complement to the subspace consisting of: (a) all polynomials in x; (b) all polynomials in x 2 ; (c) all polynomials in x with the free term equal to 0; (d) all polynomials in x with the sum of coefficients equal to 0. Problem 6. If M and N are subspaces of Hilbert space that are orthogonal to each other, then the sum M + N = {x + y : x M, y N } is a subspace. Show that the theorem is not true if M and N are either: closed but not orthogonal or orthogonal but not closed. 1.4. Weak topology Read Royden, page 236 238. Example 1.4.1. Consider the sequence of functions {cos nt} n N in L 1 [0, 2π]. It is easy to see that this sequence is not convergent. However, for any function f L, 1 0 f(t) cos nt dt 0 as n. Since L is the dual space of L 1, we say that cos nt 0 weakly, and we write w lim n cos nt = 0. Example 1.4.2. Consider the sequence of functions {cos nt} n N in L [0, 2π]. Notice that, while not a convergent sequence, if f L 1 then 1 0 f(t) cos nt dt 0 as n. Since L is the dual space of L 1, we say that cos nt 0 in the weak topology. In a Banach space X it is useful to consider three topologies: the norm topology, induced by the norm; weak topology the smallest topology in which all bounded linear functionals on X are continuous; weak topology (meaningful when X is the dual space of Y so that Y X ) the smallest topology in which some bounded linear functionals on X are continuous (those that can be identified as elements of Y). In order to dicuss these topologies (and understand their role), we need to find out what bounded linear functionals on Hilbert space H look like.

10 1. HILBERT SPACE Theorem 1.4.1 (Riesz Representation Theorem). If L is a bounded linear functional on H, then there is a unique vector y H such that L(x) = x, y for every x H. Moreover, L = y. Proof. Assuming that such y exists, we can write it as y = i N y ie i relative to a fixed o.n.b. {e i } i N. Then y i = y, e i = e i, y = L(e i ). Therefore, we define y = i N L(e i) e i, and all it remains to prove is the convergence of the series. Let s n = n i=1 L(e i) e i. Then L(s n ) = n i=1 L(e i) Le i = s n 2, so s n 2 L s n from which it follows that s n L. Lemma 1.2.5. Thus the series n i=1 L(e i) e i converges and the result follows from We see that if L H, the dual space of H, then L = L y. The mapping Φ : H H defined by Φ(y) = L y is a norm preserving surjection. It is conjugate linear: Φ(α 1 y 1 + α 2 y 2 ) = α 1 y 1 + α 2 y 2. Nevertheless, we identify H with H. Consequently, H is reflexive (i.e., H = H) so the weak and weak topologies on H coincide. Therefore, we will work with 2 topologies: weak and norm induced. The absence of a qualifier will always mean that it is the latter. Exercise 1.4.1. Prove that the weak topology is weaker than the norm toplogy, i.e., if G is a weakly open set then G is an open set. Example 1.4.3. If {e n } n N is an orthonormal sequence in H then w lim e n = 0 but the sequence is not convergent. Exercise 1.4.2. Prove that the Hilbert space norm is continuous but not weakly continuous. The following result shows why weak topology is important. [See Royden, p. 237] Theorem 1.4.2 (Banach-Alaoglu). The unit ball {x H : x 1} in Hilbert space H is weakly compact. Remark 1.4.1. The unit ball B 1 of H is NOT compact (assuming that H is infinite dimensional). Reason: if {e n } n N is an o.n.b. then the set {e 1, e 2, e 3,... } is closed but not totally bounded, hence not compact. Exercise 1.4.3. Prove that if a bounded set in H is weakly closed then it is weakly compact.

1.4. WEAK TOPOLOGY 11 In spite of the fact that the weak topology is weaker then the norm topology, some of the standard results remain true. Theorem 1.4.3. A weakly convergent sequence is bounded. Proof. Suppose that x n is a weakly convergent sequence. Then, for any y H, the sequence x n, y is a convergent sequence of complex numbers, which implies that it is bounded. In other words, for any y H there exists C = C(y) > 0 such that x n, y C. This means that, for each n N, x n can be viewed as a bounded linear functional on H. By the Uniform Bounded Principle (Royden, p. 232), these functionals are uniformly bounded, i.e., there exists M > 0 such that, for all n N, x n M. Although weakly convergent sequence need not be convergent there are situation when it does. Theorem 1.4.4. If {x n } n N is a weakly convergent sequence in a compact set K then it is convergent. Proof. Since {x n } n N K, it has an accumulation point x and a subsequence x n converging to z. If {x n } had another accumulation point x, then there would be another subsequence x n converging to w. It would follow that w lim x n = x and w lim x n = x. Since {x n } is weakly convergent this implies that x = x, so it has only one accumulation point, namely the limit. By definition, the weak topology W is the smallest one in which every bounded linear functional L on H is continuous. This means that, for any such L and any open set G in the complex plane, L 1 (G) W. Since open disks form a base of the usual topology in C it suffices to require that L 1 (G) W for each open disk G. Notice that x L 1 (G) iff L(x) G, so if G = {z : z z 0 < r} and z 0 = L(x 0 ) then x L 1 (G) iff L(x x 0 ) < r. Now Riesz Representation Theorem implies that L 1 (G) = {x H : x x 0, y < r} for some y H. We conclude that a subbase of W consists of the sets W = W (x 0 ; y, r) = {x H : x x 0, y < r}. Exercise 1.4.4. Prove that a bounded linear functional L is continuous in a topology T iff L 1 (G) T for every open disk G.

12 1. HILBERT SPACE Problem 7. Prove that a subspace of Hilbert space is closed iff it is weakly closed. Problem 8. Prove that Hilbert space is weakly complete. Problem 9. Let {x n } n N be a sequence in Hilbert space with the property that x n = 1, for all n, and x m, x n = c, if m n. Prove that {x n } n N is weakly convergent. Problem 10. Find the weak closure of the unit sphere in Hilbert space.

CHAPTER 2 Operators on Hilbert Space Nobody, except topologists, is interested in problems about Hilbert space; the people who work in Hilbert space are interested in problems about operators. Paul Halmos 2.1. Definition and Examples Read Section 10.2 in Royden s book. Operator always means linear and bounded. The algebra of all bounded linear operators on H is denoted by L(H). Example 2.1.1. Let H = C n and A = [a ij ] an n n matrix. The operator of multiplication by A is linear and bounded. Indeed, for x = (x 1, x 2,..., x n ) and M = sup 1 i n ( n j=1 a ij 2 ) 1/2, so A M. Ax = sup 1 i n n j=1 a ij x j sup 1 i n n a ij 2 j=1 1/2 n x j 2 j=1 1/2 = M x Example 2.1.2. Let H = l 2 and A = [a ij ] i,j=1, where a ij = c i if i = j and a ij = 0 if i j. We call such matrix diagonal and denote it by diag(c 1, c 2,... ), or diag(c n ). The operator A (or, more precisely, the operator of multiplication by A) is bounded iff c = (c 1, c 2,... ) l (i.e., when c is a bounded sequence). Indeed, let x = (x 1, x 2,... ) l 2, so Ax = (c 1 x 1, c 2 x 2,... ) and Ax 2 = i=1 c ix i 2. If c i M, i N, then Ax 2 M 2 i=1 x i 2 = x 2 so A is bounded. On the other hand, if c / l, then for each n there exists i n so that c in n. Then Ae in = c in e in n and A is unbounded. Remark 2.1.1. It is extremely hard to decide, in general, whether an operator A is bounded just by studying its matrix [ Ae j, e i ] i,j=1. 13

14 2. OPERATORS ON HILBERT SPACE Example 2.1.3. Let H = l 2 and let S be the unilateral shift, defined by S(x 1, x 2,... ) = (0, x 1, x 2,... ). Notice that S(x 1, x 2,... ) 2 = 0 2 + x 1 2 + x 2 2 + = x 2 so S = 1. In fact, S is an isometry, hence injective, but it is not surjective! Example 2.1.4 (Multiplication on L 2 ). Let h be a measurable function and define M h f, for f L 2, by (M h f)(t) = h(t)f(t). If h L (essentially bounded functions see Royden, p. 118), then M h f 2 = hf 2 h 2 f 2 = h 2 f 2 so M h is a bounded operator on L 2 and M h h. On the other hand, for ɛ > 0, there exists a set C [0, 1] of positive measure so that h(t) h ɛ for t C. If f = χ C then M h f 2 = hf 2 = h 2 ( h ɛ) 2 µ(c) = ( h ɛ) 2 f 2, C and it follows that M h h ɛ. We conclude that M h = h and M h is bounded iff h L. Example 2.1.5 (Integral operators on L 2 ). Let K : [0, 1] [0, 1] C be measurable and square integrable with respect to planar Lebesgue measure. We define the operator T K by (T k f)(x) = 1 K(x, y)f(y) dy. Now 0 T K f 2 = 1 0 1 0 T k f(x) 2 dx = 1 0 1 { 1 } { 1 K(x, y) 2 dy Therefore, T K is bounded and T K 0 0 K(x, y)f(y) dy 0 } f(y) 2 dy { 1 1 0 0 K(x, y) 2 dydx 2 dx 1 0 dx = f 2 1 } 1/2. ( 1 2 K(x, y)f(y) dy) dx 0 0 1 0 K(x, y) 2 dydx. Example 2.1.6 (Weighted shifts). Let H = l 2 and let {c n } n N be a bounded sequence of complex numbers. A weighted shift W on l 2 is defined by W (x 1, x 2,... ) = (0, c 1 x 1, c 2 x 2,... ). It can be written as W = S diag(c n ) so it is a bounded operator and W = diag(c n ). In some situations it is useful to have an alternate formula for the operator norm. In what follows we will use notation B 1 for the closed unit ball of H, i.e. B 1 = {x H : x 1}. Proposition 2.1.1. Let T be linear operator on Hilbert space. Then T = sup{ T x, y : x, y B 1 }.

2.2. ADJOINT 15 Proof. Let α denote the supremum above, and let us assume that T 0 (otherwise there is nothing to prove). Clearly, for x, y B 1, T x, y T, so α T. In the other direction, α sup{ T x, y : x, y B 1, T x 0, y = = sup{ T x, T x T x : x B 1, T x 0} = sup{ T x : x B 1, T x 0} T x T x } = T, and the proof is complete. 2.2. Adjoint In Linear Algebra we learn that the column space of matrix A = [a ij ] n i,j=1 and the null space of its transpose A T are orthogonal complements in R n. In C n, A T needs to be replaced by A = [a ji ] n i,j=1. In this situation, (2.1) Ax, y = x, A y. Exercise 2.2.1. Prove that, if A is an n n matrix and x, y C n, then Ax, y = x, A y. Example 2.2.1. Let h L and let M h be the operator of multiplication on L 2. Then (M h ) = M h. The following result will show that a relation (2.1) is available for any operator. Proposition 2.2.1. If T is an operator on H then there exists a unique operator S on H such that T x, y = x, Sy, for all x, y H. Proof. Let y H be fixed. Then ϕ(x) = T x, y is a bounded linear functional on H. By Riesz Representation Theorem there exists a unique z H such that ϕ(x) = x, z, for all x H. Define Sy = z. Then

16 2. OPERATORS ON HILBERT SPACE T x, y = x, Sy. To show that S is linear, let Sy 1 = z 1, Sy 2 = z 2, and let x H. Then x, S(α 1 y 1 + α 2 y 2 ) = T x, α 1 y 1 + α 2 y 2 = α 1 T x, y 1 + α 2 T x, y 2 = α 1 x, Sy 1 + α 2 x, Sy 2 = x, α 1 Sy 1 + α 2 Sy 2. By the uniqueness part of Riesz Representation Theorem S is linear. That S is unique can be deduced by contradiction: if x, Sy = x, S y for all x, y H then x, Sy S y = 0 for all x which implies that Sy S y = 0 for all y, hence S = S. Finally, S is bounded: Sy 2 = Sy, Sy = T Sy, y T Sy y T Sy y so Sy T y and S T. Definition 2.2.1. If T L(H) then the adjoint of T, denoted T, is the unique operator on H satisfying T x, y = x, T y, for all x, y H. Here are some of the basic properties of the involution T T. Proposition 2.2.2. (a) I = I (b) T = (T ) = T ; (c) T = T ; (d) (α 1 T 1 + α 2 T 2 ) = α 1 T1 + α 2 T2 ; (e) (T 1 T 2 ) = T2 T1 ; (f) if T is invertible then so is T and (T ) 1 = (T 1 ) ; (g) T 2 = T T. Proof. The assertion (a) is obvious and (b) follows from x, T y = T x, y = y, T x = T y, x = x, T y. It was shown in the proof of Proposition 2.2.1 that T T so T T T and (c) follows from (b). We leave (d) as an exercise and notice that x, (T 1 T 2 ) y = T 1 T 2 x, y = T 2 x, (T 1 ) y = x, (T 2 ) (T 1 ) y establishes (e). As a consequence of (a) and (e), T (T 1 ) = (T 1 T ) = I and (T 1 ) T = (T T 1 ) = I which

2.3. OPERATOR TOPOLOGIES 17 is (f). Finally, T T T T = T 2 and to prove the opposite inequality let ɛ > 0 and let x be a unit vector such that T x T ɛ. Then T T T T x T T x, x = T x 2 > ( T ɛ) 2, and (g) is proved. Example 2.2.2. Let H = l 2 and the let S be the unilateral shift (see Example 2.1.3). Then S (x 1, x 2,... ) = (x 2, x 3,... ). The operator S is called the backward shift. Example 2.2.3. Let T K be the integral operator on L 2 (see Example 2.1.5). Then (T K ) = T K, where K (x, y) = K(y, x). We now give the Hilbert space formulation of the relation with which we have opened this section. Theorem 2.2.3. If T is an operator on Hilbert space H then Ker T = (Ran T ). Proof. Let x Ker T and let y Ran T. Then there exists z H such that y = T z. Therefore x, y = x, T z = T x, z = 0 so x (Ran T ). In the other direction, if x (Ran T ) and z H, then T x, z = x, T z = 0. Taking z = T x we see that T x = 0, and the proof is complete. We notice that, for T L(H) and x, y H, the expression T x, y is a form that is linear in the first and conjugate linear in the second argument. It turns out that this is sufficient for a polarization identity. Proposition 2.2.4 (Second Polarization Identity). 4 T x, y = T (x + y), x + y T (x y), x y + i T (x + iy), x + iy i T (x iy), x iy. Exercise 2.2.2. Prove Second Polarization Identity. 2.3. Operator topologies In this section we take a look at the algebra L(H). It has three useful topologies which lead to 3 different types of convergence. Definition 2.3.1. A sequence of operators T n L(H) converges uniformly (or in norm) to an operator T if T n T 0, n. A sequence of operators T n L(H) converges strongly to an operator T if T n x T x 0,

18 2. OPERATORS ON HILBERT SPACE n, for all x H. A sequence of operators T n L(H) converges weakly to an operator T if T n x T x, y 0, n, for any x, y H. It follows from the definition that the weak topology is the weakest of the three, while then norm topology (a.k.a. the uniform topology) is the strongest. Are they different? Proposition 2.3.1. The operator norm is continuous with respect to the uniform topology but discontinuous with respect to the strong and weak topologies. Proof. The first assertion is a consequence of the inequality A B A B. To prove the other two, let {e n } n N be an o.n.b. of H, H n = k=n e k, P n = P Hn. Then P n 0 strongly (hence weakly) since P n x 2 = k=n+1 x k 2 0. However, P n = 1 which does not converge to 0. Example 2.3.1. We say that an operator T is a rank one operator if there exist u, v H so that T x = x, v u. We use the notation T = u v. Let T n = e n e 1. Then T n x, y = x 1 y n 0 while T n x = x 1 e n is not a convergent sequence. Thus, the weak and strong topologies are different. Example 2.3.2. The involution T T is continuous in uniform topology. ( T n T = T n T ). Also, it is continuous in the weak topology, because (T n T )x, y = x, (T n T )y = (T n T )y, x. However, it is not continuous in the strong topologies. Counterexample: let S be the unilateral shift, and T n = (S ) n. Then T n 0 strongly but {T n} is not a strongly convergent sequence. Indeed, for any x = (x 1, x 2,... ) H, T n x 2 = (x n+1, x n+2,... ) 2 = k=n x k 2 0, as n. On the other hand, for x = e 1, T nx = S n e 1 = e n, which is not a convergent sequence. An operator T L(H) is a continuous mapping when H is given the strong topology. We will write, following Halmos, (s s). One may ask about the other types of continuity. Theorem 2.3.2. The three types of continuity (s s), (w w), and (s w) are all equivalent.

2.3. OPERATOR TOPOLOGIES 19 Proof. Suppose that T is continuous, and let W be a weakly open neighborhood of T x 0 in H. We will show that T 1 (W ) is weakly open. It suffices to prove this assertion in the case when W belongs to the subbase of the weak topology. To that end, let W = W (T x 0, y, r) = {x H : x T x 0, y < r}. Then z T 1 (W ) T z W T z T x 0, y < ɛ z x 0, T y < ɛ. We see that z T 1 (W ) iff z V (x 0, T y, ɛ) so T 1 (W ) = V which is a weakly open set. The implication (w w) (s w) is trivial, so we concentrate on the implication (s w) (s s). To that end, suppose that T is not continuous. Then it is unbounded, so there exists a sequence {x n } n N of unit vectors such that T x n n 2, n N. Clearly, x n /n 0 and the assumption (s w) implies that T x n /n weakly converges to 0. By Theorem 1.4.3 the sequence {T x n /n} is bounded which contradicts the fact that T x n /n n. The fact that every operator in L(H) is weakly continuous has an interesting consequence. Corollary 2.3.3. If T is a linear operator on H then T (B 1 ) is closed. Proof. Banach-Alaoglu Theorem established that B 1 is weakly compact so, by Theorem 2.3.2, T (B 1 ) is weakly compact, hence weakly closed, hence norm closed. Exercise 2.3.1. Prove that if F is a closed and bounded set in H then T (F ) is closed. At the end of this section we consider a situation that occurs quite frequently. Theorem 2.3.4. Let M be a linear manifold that is dense in Hilbert space H. Every bounded linear transformation T : M H can be uniquely extended to a bounded linear transformation ˆT : H H. In addition, the operator norm of T equals ˆT. Proof. Let x H. Then there exists a sequence {x n } n N M converging to x. Since {x n } n N is also a Cauchy sequence, for every ɛ > 0 there exists N N such that, m, n N x m x n < ɛ/ T. It follows that, for m, n N, T x m T x n < ɛ, so {T x n } n N is a Cauchy sequence, hence convergent, and there exists y = lim n T x n. We will define ˆT x = y, i.e., ˆT (lim x n ) = lim T x n.

20 2. OPERATORS ON HILBERT SPACE First we need to establish that the definition is independent of the sequence {x n } n N. If {x n} n N is another sequence converging to x, we form the sequence (x 1, x 1, x 2, x 2,... ) which also converges to x. By the previous, the sequence (T x 1, T x 1, T x 2, T x 2,... ) must converge, and therefore, both of the subsequences {T x n } n N and {T x n} n N must have the same limit. Notice that, if x n x, the continuity of the norm implies that ˆT x = lim T x n = lim T x n lim T x n = T x so ˆT T. Since the other inequality is obvious we see that ˆT = T. In particular, ˆT is a bounded operator. Also, ˆT (αx + βy) = ˆT (α lim x n + β lim y n ) = ˆT (lim(αx n + βy n )) = lim T (αx n + βy n ) = lim(αt x n + βt y n ) = α lim T x n + β lim T y n = α ˆT x + β ˆT y, so ˆT is linear. Finally, suppose that T 1 and T 2 are two continuous extensions of T, and let x H. If x n x, the continuity implies that both T 1 x n T 1 x and T 2 x n T 2 x. If x n M then T 1 x n = T 2 x n, so T 1 x = T 2 x. Therefore, the extension is unique, and the proof is complete. Need an example 2.4. Invariant and Reducing Subspaces When M is a closed subspace of H, we can always write H = M M. Relative to this decomposition, any operator T acting on H can be written as a 2 2 matrix with operator entries (2.2) T = X Z Y. W It is sometimes convenient to consider only the initial space or the target space as a direct sum. In such a situation [ ] we will use a 1 2 or 2 1 matrix. Thus X Y will describe an operator T : M M H; if f M and g M then [ X Y ] [ ] f g = Xf + Y g. A subspace M is invariant for T if, for any x M, T x M. It is reducing for T if both M and M are invariant for T.

2.4. INVARIANT AND REDUCING SUBSPACES 21 Example 2.4.1. The subspace (0) consisting of zero vector only is an invariant subspace for any operator T. Also, H is an invariant subspace for any operator T. Because they are invariant for every operator they are called trivial. A big open problem in Operator theory is whether every operator has a non-trivial invariant subspace. Example 2.4.2. If M is a closed subspace of H and T 1 is an operator on M with values in M, then the operator T = T 1 0, defined by T x = T 1 x if x M and T x = 0 if x M is an operator in L(H). However, if M is not invariant for T 1, the same definition (T x = T 1 x for x M, T x = 0 for x M ) describes the operator [ ] T 1 0. Proposition 2.4.1. If T is an operator on Hilbert space H, and P = P M is the projection onto the closed subspace M, then the following are equivalent: (a) M is invariant for T ; (b) P T P = T P ; (c) Z = 0 in (2.2). Proof. It is not hard to see that the matrix for P is [ I 0 0 0 ] so P T P T P = [ ] 0 0 Z 0. This establishes (b) (c). Since [ ] [ f g M iff g = 0, we see that T f ] [ ] 0 = Xf Zf M for all x H iff Z = 0 so (a) (c). Example 2.4.3. Let S be the unilateral shift, n N, and M = k n e k. Then SM = k n+1 e k M. Proposition 2.4.2. If T is an operator on Hilbert space H, and P = P M then the following are equivalent: (a) M is reducing for T ; (b) P T = T P ; (c) Y, Z = 0 in (2.2); (d) M is invariant for T and T. Proof. Since P T T P = [ 0 Y Z 0 ] we see that (b) (c). Further, the matrix for T is [ X Z Y W ] so, by Proposition 2.4.1, M is invariant for T and T iff Z = Y = 0 and (c) (d). In order to prove that (a) (d)

22 2. OPERATORS ON HILBERT SPACE it suffices to show that M is invariant for T iff M is invariant for T. By Proposition 2.4.1, M is invariant for T iff Y = 0 (iff Y = 0). On the other hand T [ ] [ ] 0 g = Y g W g M iff Y g = 0 for all g. Exercise 2.4.1. Prove that the matrix for T is [ ] X Z Y W. Example 2.4.4. Let T = M h, let E [0, 1], m(e) > 0, and let M = L 2 (E). If f M then T f = hf M. Also, T = M h and T f = hf M, so M is reducing for T. Example 2.4.5. Let S be the unilateral shift, n N, and M = k n e k. Then M is invariant for S but not reducing, since e n M but S e n = e n 1 / M. 2.5. Finite rank operators The closest relatives of finite matrices are the finite rank operators. Definition 2.5.1. An operator T is a finite rank operator if its range is finite dimensional. We denote the set of finite rank operators by F. Example 2.5.1. If T is a rank one operator u v (see Example 2.3.1) then the range of u v is the one dimensional subspace spanned by u, so u v F. The rank one operators turn out to be the building blocks out of which finite rank operators are made. Proposition 2.5.1. If T is a linear operator on H then T belongs to F iff there exist vectors u 1,u 2,...,u n, and v 1,v 2,...,v n such that T x = n i=1 x, v i u i. Proof. Suppose that Ran T is of finite dimension n, and let e 1, e 2,..., e n be an o.n.b. of Ran T. Then T x = n i=1 T x, e i e i = n i=1 x, T e i e i. We leave the converse as an exercise. Exercise 2.5.1. Prove that if there exist vectors u 1, u 2,..., u n, v 1, v 2,..., v n such that T x = n i=1 x, v i u i, for all x H, then Ran T is of dimension at most n.

2.6. COMPACT OPERATORS 23 Exercise 2.5.2. Prove that if T = u i v i then T = v i u i. The next theorem summarizes some very important properties of the class F. Theorem 2.5.2. The set F is a minimal -ideal in L(H). Here the star means that F is closed under the operation of taking adjoints. Proof. It is obvious that F is a subspace of L(H). Furthermore, if T F and A L(H), then Ran T A Ran T so T A F. Also, if T is of finite rank, then according to Proposition 2.5.1, T = n i=1 u i v i so T = n i=1 v i u i. It follows that T F and the same is true of T A, for any A L(H). Consequently, AT is of finite rank, and F is a -ideal. To see that it is minimal, it suffices to show that, if J is a non-zero ideal, then J contains all rank one operators. Let T J, T 0. Then there exists vectors x, y, such that y = 1 and y = T x. Let u v be a rank one operator. Since J is an ideal, it contains the product (u y)t (x v) which equals u v. A finite rank operator is a generalization of a finite matrix. What happens when we take the closure of F in some topology? Exercise 2.5.3. Prove that the strong closure of F is L(H). [Hint: Prove that P n I strongly.] Conclude that the weak closure of F is also L(H). 2.6. Compact Operators Exercise 2.5.3 established that the strong closure of F is L(H). Therefore, we consider the norm topology. Definition 2.6.1. An operator T in L(H) is compact if it is the limit of a sequence of finite rank operators. We denote the set of compact operators by K. Example 2.6.1. Let T = diag(c n ) as in Example 2.1.2, with lim n c n = 0. Then T is compact. Reason: take T n = diag(c 1, c 2,..., c n, 0, 0,... ). Then T n F and T T n = sup{ c k : k n + 1} 0. It follows that T is compact.

24 2. OPERATORS ON HILBERT SPACE Example 2.6.2. Let T = T K as in Example 2.1.5. If K L 2 ([0, 1] [0, 1]) then T K is compact. We will point out at several different sequences in F that all converge to T K We start with a function theoretic approach: simple functions are dense in L 2 (Royden, p. 128), and a similar proof establishes that simple functions are dense in L 2 ([0, 1] [0, 1]). Since a simple function is a linear combination of the characteristic functions of rectangles χ [a,b] [c,d] (x, y) = χ [a,b] (x)χ [c,d] (y) it follows that K(x, y) is the L 2 limit of functions of the form K n (x, y) = n i=1 f i(x)g i (y), so T K is the norm limit of T Kn, which are all finite rank operators. Exercise 2.6.1. Verify that T Kn F, if K n (x, y) = n i=1 f i(x)g i (y). Our second approach is exploiting the fact that L 2 is Hilbert space. If {e j } j N is an o.n.b. of L 2 we can, for a fixed y, write K(x, y) = j=1 k j(y)e j (x). Now define K N (x, y) = N j=1 k j(y)e j (x) and notice that T KN T K as N. Exercise 2.6.2. Verify that T KN F and that lim N T KN = T K, if K N (x, y) is as above. Our last method is based on the matrix for T K. Let k ij = T K e j, e i, with {e n } n N an o.n.b. of L 2 ([0, 1]). First we notice that, for any f L 2, k f, e k 2 = k f, e k e k 2 = f 2. Therefore, T K e j, e i 2 = e j, TKe i 2 = TKe i, e j, 2 = TKe i 2 j=1 j=1 j=1 = 1 It follows that, for any n N, n k ij 2 = 0 1 i=1 j=1 i=1 0 0 K (y, x)e i (x) dx n 1 1 0 1 1 0 i=1 0 2 dy = K(x, y)e i (x) dx K(x, y)e i (x) dx 2 2 1 0 1 0 dy = K(x, y)e i (x) dx 1 n 1 0 i=1 0 1 1 2 dy. K(x, y)e i (x) dx dy = K(x, y) 2 dxdy 0 0 2 dy

2.6. COMPACT OPERATORS 25 so the series k ij 2 converges. Operators whose matrices satisfy this condition are called Hilbert Schmidt i=1 j=1 operators. The Hilbert-Scmidt norm is defined as T K 2 = { i=1 j=1 k ij 2 } 1/2, and it satisfies the inequality A A 2. Hilbert-Scmidt operators are compact because we can define T n to be the matrix consisting of the first n rows of the matrix of T K and having the remaining entries 0. Then each T n F and T n T K 0. Indeed, Ran T n {e 1, e 2,..., e n }, and T K T n 2 T K T n 2 2 = k ij 2 0, n. i=n+1 j=1 Exercise 2.6.3. Prove that the Hilbert-Scmidt norm is indeed a norm and, for any T L(H), T T 2. Next we consider some of the properties of compact operators. The first one follows directly from the definition. Theorem 2.6.1. The set K is the smallest closed -ideal in L(H). The following result reveals the motivation for calling these operators compact. Theorem 2.6.2. An operator T in L(H) is compact iff it maps the closed unit ball of H into a compact set. Proof. Suppose that K is compact and let {y n } n N be a sequence in K(B 1 ). We will show that there exists a subsequence of {y n } that converges to an element of K(B 1 ). Notice that, for every n N, y n = Kx n, and x n belongs to the weakly compact set B 1. Thus, there exists a subsequence {x nk } converging weakly to x B 1. Thus, it suffices to show that Kx nk converges to Kx. Let {K n } be a sequence in F thaty converges to K. For any m N, K m (B 1 ) is a bounded and closed set (by Corollary 2.3.3) that is contained in a finite dimensional subspace of H, so it is compact. By Theorem 1.4.4, {K m x nk } k N converges to K m x. Now, let ɛ > 0. Then there exists N N such that K K N < ɛ/3. Further, with N fixed, there exists k 0 N so that, for k k 0, K N x nk K N x < ɛ/3. Therefore, for k k 0, Kx nk Kx (K K N )x nk + K N (x nk x) + (K N K)x < ɛ 3 + ɛ 3 + ɛ 3 = ɛ. Thus, y nk = Kx nk is a convergent subsequence converging to Kx K(B 1 ) so K(B 1 ) is a compact set.

26 2. OPERATORS ON HILBERT SPACE Suppose now that K(B 1 ) is compact and let n N. Notice that y K(B1)B(y, 1/n) is an open covering of the compact set K(B 1 ), so there exist vectors x (n) 1, x(n) 2,..., x(n) k H so that k i=1 B(Kx(n) i, 1/n) is a covering of K(B 1 ). Let H n be the span of Kx (n) 1, Kx(n) 2,..., Kx(n) k and P n the orthogonal projection on H n. Finally, let K n = P n K. Clearly, K n F. Let ɛ > 0, and choose N > 1/ɛ. If n N, and x 1, then Kx K n x = Kx P n Kx. Since P n Kx is the point in H n closest to Kx, it follows that Kx K n x inf 1 i n Kx Kx (n) i < 1/n < ɛ. Thus K n K and the proof is complete. Remark 2.6.1. In many texts the characterization of compact operators, established in Theorem 2.6.2, is taken to be the definition of a compact operator. Exercise 2.6.4. Prove that if F is a closed and bounded set in H and T is a compact operator in L(H) then T (F ) is a compact set. There is another characterization of compact operators: Proposition 2.6.3. If T is a linear operator on H then T is compact iff it maps every weakly convergent sequence into a convergent sequence. In this situation, if w lim x n = x then lim T x n = T x. Proof. Suppose first that T is compact and let w lim x n = x. By Proposition 1.4.3, there exists M > 0 such that, for all n N, x n M. Therefore, T x n /M T (B 1 ), which is compact by Theorem 2.6.2. Now Theorem 1.4.4 implies that lim x n = x. In order to establish the converse, we will demonstrate that T (B 1 ) is compact by showing that every sequence in T (B 1 ) has a convergent subsequence. Let {y n } n N T (B 1 ). Then y n = T x n, for x n B 1, so the Banach Alaoglu Theorem implies that {x n } has a weakly convergent subsequence {x nk } and, by assumption, {T x nk } is a (strongly) convergent subsequence of {T x n }. Example 2.6.3. We have seen in Example 2.6.1 that if T diag(c n ) and c n 0, then T is compact. The converse is also true: if {e n } is the o.n.b. which makes T diagonal, then T e n 0 (because w lim e n = 0 and T is compact) so c n e n 0.

2.7. NORMAL OPERATORS 27 It is useful to know that compactness is inherited by the parts of an operator. Theorem 2.6.4. Suppose that T is a compact operator on Hilbert space H = M M and that, relative to this decomposition, T = [ X Z W Y ]. Then each of the operators X, Y, Z, W is compact. Proof. Let {T n } be a sequence of finite rank operators that converges to T. Write, for each n N, Y n T n = [ X n ] Z n W n. Then all the operators Xn, Y n, Z n, W n F and they converge to X, Y, Z, W, respectively. Exercise 2.6.5. Prove that X n, Y n, Z n, W n F and that they converge to X, Y, Z, W, respectively. [Consider the projections P 1 = P M and P 2 = P M and notice that, for example P 1 T P 2 = [ 0 Y 0 0 ], so Y n Y T n T and Ran Y n Ran P 1 T n P 2 the later being finite dimensional.] 2.7. Normal operators Definition 2.7.1. If T is an operator on Hilbert space H then: (a) T is normal if T T = T T ; (b) T is self-adjoint (or Hermitian) if T = T ; (c) T is positive if T x, x 0 for all x H; (d) T is unitary if T T = T T = I. Example 2.7.1. Let T = diag(c n ). Then T = diag(c n ) so T is normal. Also, T = T iff c n R, n N, and T is positive iff c n 0, n N. Finally, T T = diag( c n 2 ) so T is unitary iff c n = 1, n N. Exercise 2.7.1. Let T = M h on L 2. Prove that T is normal and that it is: self-adjoint iff h(x) R, a.e.; positive iff h(x) 0 a.e.; unitary iff h(x) = 1 a.e.. The relationship between T and T that defines each of these classes allows us to establish some of their significant properties. Proposition 2.7.1. An operator T on Hilbert space H is self-adjoint iff T x, x is real for any x H.

28 2. OPERATORS ON HILBERT SPACE Proof. If T = T then T x, x = x, T x = x, T x = T x, x so T x, x R. On the other hand, if T x, x is real for any x H then Second Polarization Identity implies that T x, y = T y, x so T = T. Exercise 2.7.2. Prove that T x, x R implies that T x, y = T y, x. Corollary 2.7.2. If P is a positive operator on Hilbert space H then P is self-adjoint. Example 2.7.2. If P is the orthogonal projection on a subspace M of Hilbert space H, then P is a positive operator. Indeed, if z H write z = x + y relative to H = M M. By Theorem 1.3.2, P z = x and P y = 0, so P z, z = x, x + y = x 2 0. Combining Theorem 1.3.2 and Example 2.7.2 we see that every projection is a positive idempotent. In fact, the converse is also true. Theorem 2.7.3. If T is an idempotent self-adjoint operator then T is a projection on M = {x H : T x = x}. Proof. Let z H and write it as z = T z + (z T z). Now T (T z) = T z so T z M. Also, z T z M. Indeed, if x M, then x, z T z = x, z x, T z = x, z T x, z = 0. By Proposition 2.1.1, the norm of every operator T in L(H) can be computed by considering the supremum of the values of its bilinear form T x, y. The next result shows that, when T is self adjoint, it suffices to consider only some pairs of x, y B 1. Proposition 2.7.4. If T is a self-adjoint operator on Hilbert space H then T = sup{ T x, x : x = 1}. Proof. Clearly, T x, x T x 2, so if we denote by α the supremum above, we have that α T. To prove that α = T, we use the Second Polarization Identity, and we notice that, in view of the assumption T = T and Proposition 2.7.1, 4Re T x, y = T (x + y), x+y T (x y), x y. Moreover, using Parallelogram Law, and assuming that x and y are unit vectors, we obtain that 4Re T x, y α x+y 2 +α x y 2 = α(2 x 2 +2 y 2 ) = 4α. When x is selected so that T x 0, y = T x/ T x we obtain Re T x α so T α.

2.7. NORMAL OPERATORS 29 Exercise 2.7.3. Prove that two product of two self-adjoint operators is self-adjoint iff the operators commute. Remark 2.7.1. If we write A = (T + T )/2 and B = (T T )/2i then the operators A, B are self-adjoint and T = A + ib. We call them the real part and the imaginary part of T. Proposition 2.7.5. If T is an operator on Hilbert space H then the following are equivalent. (a) T is a normal operator; (b) T x = T x for all x H; (c) the real and imaginary part of T commute. Proof. Notice that T x 2 T x 2 = (T T T T )x, x. If T is normal then the right side is 0, so (a) implies (b). If (b) is true, then the left side is 0, for all x. Since T T T T is self-adjoint, Proposition 2.7.4 implies that its norm is 0, so (b) implies (a). A calculation shows that, if A and B are the real and imaginary part of T, resp., then AB BA = (T T T T )/2i so (a) is equivalent to (c). In Definition 1.2.3 we have introduced the concept of the Hilbert space isomorphsim. Since it preserves the inner product ( Ux, Uy = x, y ), it preserves the norm, and hence both weak and strong toplogies. Therefore, if U : H K, we do not distinguish between an operator T L(H) and UT U 1 L(H), and we say that they are unitarily equivalent. Since, by Definition 2.7.1, an operator T is unitary iff T T = T T = I, we should check that UU = U U = I. Exercise 2.7.4. Verify that UU = I K and U U = I H. Notice that both equalities need to be verified, because it is quite possible for one to hold but not the other. Example: the unilateral shift S satisfies S S = I SS. Exercise 2.7.5. Prove that T is an isometry iff T T = I. Exercise 2.7.3 asserts that the product of two self-adjoint operators is itself self-adjoint iff the operators commute. What if self-adjoint is replaced by normal? If M, N are commuting normal operators, their product

30 2. OPERATORS ON HILBERT SPACE is normal if MN commutes with N M and it looks like we need the additional assumption that M commutes with N (which also gives that M commutes with N). When an operator T commutes with both N and N we say that T doubly commutes with N. When N is normal we can establish even a stronger result. Theorem 2.7.6 (Fuglede Putnam Theorem). Suppose that M, N are normal operators and T L(H) intertwines M and N: MT = T N. Then M T = T N. Proof. Let λ be a complex number, and denote A = λm, B = λn. Notice that AT = T B, so A 2 T = A(AT ) = A(T B) = (AT )B = (T B)B = T B 2, and inductively A k T = T B K, for k N. It follows that, if we denote the exponential function by exp(z), exp(a)t = T exp(b). It is not hard to see that exp( A) exp(a) = I so T = exp( A)T exp(b). If we denote by U 1 = exp(a A), U 2 = exp(b B ), then both U 1, U 2 are unitary operators. Indeed, U 1 = [ (A A) n /n!] = (A A ) n /n! = exp(a A ) = U 1 1, and similarly for U 2. Now we have that exp(a )T exp( B ) = U 1 T U 2 and exp(a )T exp( B ) = T. We conclude that exp(λm )T exp( λn ) = T for all λ C. Now f(λ) = exp(λm )T exp( λn ) is an entire bounded function, hence a constant. Therefore, f (0) = 0. On the other hand, f (λ) = M exp(λm )T exp( λn ) + exp(λm )T exp( λn )( N ) so f (0) = M T T N, and the theorem is proved. Exercise 2.7.6. Prove that exp( T ) exp(t ) = I for any operator T L(H). Corollary 2.7.7. The product of two normal operators is itself normal iff the operators commute. Exercise 2.7.7. Prove Corollary 2.7.7.